Sei sulla pagina 1di 115

Chapter 5

SUBSTRUCTURES

Smith Bridge, Houlton

West Bridge, Fairfield-Benton

CHAPTER 5 - SUBSTRUCTURES

SUBSTRUCTURES ....................................................................................5-1
5.1 Terminology .........................................................................................5-1
5.2 General ................................................................................................5-3
5.2.1
Frost ..........................................................................................5-3
5.2.2
Seal Cofferdams ........................................................................5-6
5.2.3
Cofferdams ................................................................................5-7
5.2.4
Concrete Joints ..........................................................................5-8
5.2.5
Seismic Considerations .............................................................5-9
5.3 Spread Footings ................................................................................5-10
5.3.1
Service Limit States .................................................................5-10
5.3.2
Strength Limit States ...............................................................5-11
5.3.3
Extreme Event Limit States......................................................5-12
5.3.4
Footing Depth ..........................................................................5-12
5.3.4.1 Bearing Materials ....................................................................5-12
5.3.4.2 Footings on Bedrock ...............................................................5-13
5.3.4.3 Frost Protection.......................................................................5-13
5.3.4.4 Scour Protection .....................................................................5-13
5.3.5
Bearing Resistance ..................................................................5-13
5.3.5.1 General ...................................................................................5-13
5.3.5.2 Bearing Stress Distribution .....................................................5-14
5.3.5.3 Bearing Resistance Factors ....................................................5-15
5.3.6
Settlement................................................................................5-16
5.3.6.1 Tolerable Settlement ...............................................................5-17
5.3.6.2 Settlement Analyses ...............................................................5-17
5.3.7
Overall Stability ........................................................................5-18
5.3.8
Sliding ......................................................................................5-18
5.3.9
Eccentricity ..............................................................................5-20
5.3.10
Ground Water Condition ..........................................................5-20
5.3.11
Drainage Considerations .........................................................5-20
5.4 Abutments .........................................................................................5-20
5.4.1
Conventional Abutments ..........................................................5-20
5.4.1.1 General Design Requirements ................................................5-20
5.4.1.2 Loads Combinations and Load Factors ..................................5-21
5.4.1.3 General ...................................................................................5-23
5.4.1.4 Strength Limit State Evaluations .............................................5-24
5.4.1.5 Service Limit State Evaluations ..............................................5-24
5.4.1.6 Extreme Limit State Evaluations .............................................5-25
5.4.1.7 Load Considerations ...............................................................5-25
5.4.1.8 Backfill ....................................................................................5-26
5.4.1.9 Drainage .................................................................................5-26
5.4.1.10 Reinforcement and Structural Design .....................................5-27
5.4.1.11 Abutments on Spread Footings ..............................................5-27
5.4.1.12 Abutments Supported on Pile Foundations .............................5-37
5.4.1.13 Bridge Seat Dimensions .........................................................5-38
5.4.2
Integral Abutments ...................................................................5-38
5.4.2.1 Introduction .............................................................................5-38

March 2014

CHAPTER 5 - SUBSTRUCTURES

5.4.2.2 Loads ......................................................................................5-39


5.4.2.3 Historical ASD Design Practice and Bridge Lengths ...............5-39
5.4.2.4 Pile Design ..............................................................................5-40
5.4.2.5 Pile Length Requirement ........................................................5-43
5.4.2.6 Maximum Bridge Lengths .......................................................5-45
5.4.2.7 Best Practices for Moderate to Long Span IABs ...................5-46
5.4.2.8 Abutment Details.....................................................................5-48
5.4.2.9 Alignment ................................................................................5-51
5.4.2.10 Superstructure Design ............................................................5-51
5.4.2.11 Abutment and Wingwall Design ..............................................5-51
5.4.2.12 Approach Slabs.......................................................................5-52
5.4.2.13 Drainage .................................................................................5-53
5.4.2.14 Scour ......................................................................................5-53
5.4.2.15 Integral Abutment on Spread Footing Design .........................5-53
5.4.3
Semi-Integral Abutments .........................................................5-53
5.4.4
Approach Slabs .......................................................................5-54
5.5 Piers ..................................................................................................5-55
5.5.1
Mass Piers ...............................................................................5-55
5.5.1.1 Pier Selection Criteria .............................................................5-55
5.5.1.2 Load Combinations and Load Factors ....................................5-55
5.5.1.3 General ...................................................................................5-58
5.5.1.4 Strength Limit State Evaluations .............................................5-58
5.5.1.5 Service Limit State Evaluations ..............................................5-59
5.5.1.6 Extreme Event Limit State Evaluations ...................................5-59
5.5.1.7 Structural Design ....................................................................5-60
5.5.1.8 Structural Design of Columns .................................................5-60
5.5.1.9 Geotechnical Design of Pier Foundations ...............................5-61
5.5.1.10 Pier Protection ........................................................................5-61
5.5.2
Pile Bent Piers .........................................................................5-63
5.5.2.1 Pile Bent Use Criteria..............................................................5-63
5.5.2.2 Loads and Load Combinations ...............................................5-64
5.5.2.3 Pile Cap Design ......................................................................5-67
5.5.2.4 Pile Type Selection Criteria .....................................................5-67
5.5.2.5 Pile Protection .........................................................................5-68
5.5.2.6 Pipe Pile Coatings and Cathodic Protection ...........................5-69
5.5.2.7 Additional Pile Bent Pier Design Criteria .................................5-71
5.6 Retaining Walls ..................................................................................5-75
5.6.1
General ....................................................................................5-75
5.6.1.1 Retaining Wall Type Selection ................................................5-75
5.6.1.2 Service Life .............................................................................5-76
5.6.1.3 Design Loads ..........................................................................5-76
5.6.1.4 Limit States .............................................................................5-77
5.6.1.5 Strength Limit State ................................................................5-79
5.6.1.6 Service Limit State Checks .....................................................5-80
5.6.1.7 Design Considerations ............................................................5-81
5.6.1.8 Aesthetics ...............................................................................5-81

March 2014

ii

CHAPTER 5 - SUBSTRUCTURES

5.6.2
Gravity Retaining Walls............................................................5-81
5.6.2.1 Design Section ........................................................................5-81
5.6.2.2 Earth Loads ............................................................................5-82
5.6.3
Gravity Cantilever-type Retaining Walls ..................................5-82
5.6.3.1 Design Section Gravity Cantilever Retaining Walls.................5-82
5.6.3.2 Earth Loads ............................................................................5-83
5.6.4
Non-Gravity Cantilever and Anchored Retaining Walls ...........5-84
5.6.4.1 Soil Nail Walls .........................................................................5-84
5.6.5
Prefabricated Proprietary Walls ...............................................5-84
5.6.5.1 Proprietary Retaining Walls ....................................................5-85
5.6.5.2 Prefabricated Concrete Modular Gravity Walls .......................5-85
5.6.5.3 Precast Concrete Block Gravity Walls ....................................5-87
5.6.5.4 MSE Walls ..............................................................................5-87
5.6.5.5 PS&E for Project with Proprietary Walls .................................5-91
5.6.6
Geosynthetic Reinforced Soil Integrated Bridge Systems .......5-92
5.6.7
Anchored Wall Systems ...........................................................5-92
5.6.7.1 CON/SPAN Wingwall ............................................................5-92
5.6.7.2 Metal Structural Plate Headwall/Wingwall ...............................5-93
5.6.8
Gabions ...................................................................................5-93
5.7 Piles ...................................................................................................5-94
5.7.1
General ....................................................................................5-94
5.7.2
H-Piles .....................................................................................5-94
5.7.2.1 Axial Resistance .....................................................................5-94
5.7.2.2 Lateral Pile Resistance for the Service Limit State .................5-96
5.7.3
Layout and Construction ..........................................................5-99
5.7.4
Concrete Piles .........................................................................5-99
5.7.5
Steel Pipe Piles ......................................................................5-100
5.7.5.1 Design - General ...................................................................5-100
5.7.5.2 Material and Design Section .................................................5-100
5.7.6
Downdrag ..............................................................................5-101
5.7.7
Pile Installation Quality Control and Nominal Pile Resistance5-102
5.8 Drilled Shafts ...................................................................................5-105
5.9 Embankment Issues ........................................................................5-106
5.9.1
Embankment Settlement........................................................5-106
5.9.2
Embankment Stability ............................................................5-106
5.9.3
Embankment Bearing Capacity .............................................5-108
5.9.4
Embankment Seismic Considerations ...................................5-108
References ....................................................................................................5-109
Table 5-1 Depth of Frost Penetration .................................................................5-3
Table 5-2 Bearing Resistance Factors ............................................................5-16
Table 5-3 Resistance Factors for Sliding of Spread Footings at the Strength
Limit State .................................................................................................5-19
Table 5-4 Typical Load Groups and Load Factors (i) for Abutments on Spread
Footings ....................................................................................................5-23

March 2014

iii

CHAPTER 5 - SUBSTRUCTURES

Table 5-5 Recommended Maximum Lengths for Fully Integral Abutment Bridges
(feet) .........................................................................................................5-46
Table 5-6 Typical Load Groups and Load Factors ..........................................5-79
Table 5-7 Factored Axial Structural Resistance of Selected H-Pile Sections ..5-96
Table 5-8 Factored Lateral Resistance and Depth to Fixity for Strength Limit
State Design for H-Pile Sections in Sand, =1.0 ......................................5-97
Table 5-9 Factored Lateral Resistance and Depth to Fixity for Service Limit State
Design for H-Pile Sections in Clay, =1.0, Load Perpendicular to Flange 5-98
Table 5-10 Resistance Factors for Driven Piles ............................................5-103
Figure 5-1 Maine Design Freezing Index Map ...................................................5-5
Figure 5-2 Fixed Pile Head, Full Integral Abutment Details-Steel Superstructures
..................................................................................................................5-49
Figure 5-3 Integral Abutment with Hinge and Full Integral Abutment Details
Precast Superstructures ...........................................................................5-50
Figure 5-4 Effective Pile Length for Piles in Sand .........................................5-73
Figure 5-5 Effective Pile Length for Piles in Clay ...........................................5-74
Figure 5-6 Retaining Wall Design Section .......................................................5-83
Procedure 5-1 Bearing Stress on Soil ..............................................................5-30
Procedure 5-2 Bearing Stress on Bedrock .......................................................5-32
Procedure 5-3 Eccentricity and Sliding Check for Conventional Abutment on
Spread Footing .........................................................................................5-35
Example 5-1 Depth of Frost Penetration ............................................................5-6

March 2014

iv

CHAPTER 5 - SUBSTRUCTURES

5 SUBSTRUCTURES
5.1

Terminology

B
B
C
D
DLV, LLv
Df
e
eo
Ep
Eg
F.G.
H
Ht
Ip
Ig
K
Ka
Kho
Ko
Kp
L
L
Le
Ls
Lu
Lus
M
Mo
Mr
Mt
O
Ph,q
Ph
PL
Pp
Pt

March 2014

footing width
effective footing width
point designating center of footing
height of soil in front of structure, which is applicable to passive
resistance
vertical structural/superstructure loads applied to abutment wall
depth to fixity
eccentricity of the resultant of all vertical forces at the bottom of the
footing, measured from mid-width of footing
eccentricity calculated about the toe of the footing, to be used for
overturning calculations
modulus of elasticity of pile
modulus of elasticity of end span beam/girder
finished grade elevation
height of structure or failure plane
horizontal force required to translate pile
moment of inertia of pile
moment of inertia of end span beam/girder (composite I for
composite beams)
effective length factor
active earth pressure coefficients for level or sloped backfill
active earth pressure coefficient corresponding to a broken
backslope
at-rest earth pressure coefficient
passive earth pressure coefficient.
heel length
effective footing length
effective pile length from ground surface to the point of assumed
fixity below ground, including scour effects.
length of end span
exposed pile length above ground
unsupported length
pile head moment
overturning moment
resisting moment
moment induced in the pile from the horizontal translation
point designating the toe of footing
horizontal traffic surcharge force behind abutment wall
horizontal soil active force behind abutment wall
allowable lateral load
horizontal passive force
pile reaction resulting from the earth pressure on the abutment

5-1

CHAPTER 5 - SUBSTRUCTURES

qs
Q
Qapplied
R
Rn
RR
Rf
Rg
Sp
t
w
W
Wc1, Wc2
Ws
Wtoe
XDL
XLL
XWS
XWC1
XWC2
Xwtoe
y
z

i
i
p

c
f

bc
dyn
stat
ep

March 2014

traffic live load surcharge pressure


factored horizontal sliding force
applied load or stress
resultant force at base of footing
nominal resistance of footing, pile, shaft or micropile
factored resistance of a footing, pile, shaft or micropile
factored bearing or sliding resistance of a footing
beam/girder rotation (radians)
section modulus of the pile
footing thickness
water content (percent)
total beam/girder live load, end span
weight of abutment wall, footing
weight of soil above heel
weight of soil above toe
distance from the point of interest to the dead load reaction
(centerline of bearing)
distance from the point of interest to the live load reaction
(centerline of bearing)
distance from the point of interest to the centroid of Ws
distance from the point of interest to the centroid of Wc1
distance from the point of interest to the centroid of Wc2
distance from the point of interest to the centroid of Wtoe
the depth of seal from top of seal to bottom of seal
the depth of water from water surface to bottom of seal
batter angle from the horizontal plane
backfill slope
friction angle between soil/bedrock and concrete
soil weight
column slenderness factor
factors to account for ductility, redundancy and operational
importance
load factor (general)
permanent load factor
soil internal angle of friction
factored bearing stress at base of footing
horizontal superstructure forces transmitted through bearing at wall
top
resistance factor for axial compression
resistance factor for flexure
resistance factor (general - geotechnical)
resistance factor for bearing resistance
resistance factor for driven piles, dynamic analysis methods
resistance factor for piles, static analysis methods
resistance factor for passive soil resistance
resistance factor for sliding resistance between footing and soil/rock
5-2

CHAPTER 5 - SUBSTRUCTURES

5.2

General

5.2.1 Frost
Any foundation placed on seasonally frozen soils must be embedded below
the depth of frost penetration to provide adequate frost protection and to
minimize the potential for freeze/thaw movements. Fine-grained soils with low
cohesion tend to be most frost susceptible. Soils containing a high percentage
of particles smaller than the No. 200 sieve also tend to promote frost
penetration.
In order to estimate the depth of frost penetration at a site, Table 5-1 has been
developed using the Modified Berggren equation and Figure 5-1 Maine Design
Freezing Index Map. The use of Table 5-1 assumes site specific, uniform soil
conditions where the Geotechnical Designer has evaluated subsurface
conditions. Coarse-grained soils are defined as soils with sand as the major
constituent. Fine-grained soils are those having silt and/or clay as the major
constituent. If the make-up of the soil is not easily discerned, consult the
Geotechnical Designer for assistance. In the event that specific site soil
conditions vary, the depth of frost penetration should be calculated by the
Geotechnical Designer.
Table 5-1 Depth of Frost Penetration
Design
Freezing
Index
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900
2000
2100
2200
2300
2400
2500
2600

March 2014

Frost Penetration (in)


Coarse Grained
Fine Grained
w=10% w=20% w=30% w=10% w=20% w=30%
66.3
55.0
47.5
47.1
40.7
36.9
69.8
57.8
49.8
49.6
42.7
38.7
73.1
60.4
52.0
51.9
44.7
40.5
76.3
63.0
54.3
54.2
46.6
42.2
79.2
65.5
56.4
56.3
48.5
43.9
82.1
67.9
58.4
58.3
50.2
45.4
84.8
70.2
60.3
60.2
51.9
46.9
87.5
72.4
62.2
62.2
53.5
48.4
90.1
74.5
64.0
64.0
55.1
49.8
92.6
76.6
65.7
65.8
56.7
51.1
95.1
78.7
67.5
67.6
58.2
52.5
97.6
80.7
69.2
69.3
59.7
53.8
100.0
82.6
70.8
71.0
61.1
55.1
102.3
84.5
72.4
72.7
62.5
56.4
104.6
86.4
74.0
74.3
63.9
57.6
106.9
88.2
75.6
75.9
65.2
58.8
109.1
89.9
77.1
77.5
66.5
60.0

5-3

CHAPTER 5 - SUBSTRUCTURES

Notes: 1. w = water content


2. Where the Freezing Index and/or water content is between the
presented values, linear interpretation may be used to determine
the frost penetration.

March 2014

5-4

Figure 5-1 Maine Design Freezing Index Map

March 2014

CHAPTER 5 - SUBSTRUCTURES

Example 5-1 illustrates how to use Table 5-1 and Figure 5-1 to determine the
depth of frost penetration:
Example 5-1 Depth of Frost Penetration
Given:
Site location is Freeport, Maine
Soil conditions: Silty fine to coarse Sand
Step 1. From Figure 5-1 Design Freezing Index = 1300 degree-days
Step 2. From laboratory results: soil water content = 28% and major constituent Sand
Step 3. From Table 5-1: Depth of frost penetration = 56 inches = 4.7 feet

Spread footings founded on bedrock require no minimum embedment depth.


Pile supported footings will be embedded for frost protection. The minimum
depth of embedment will be calculated using the techniques discussed in
Example 5-1. Pile supported integral abutments will be embedded no less
than 4.0 feet for frost protection.
Riprap is not to be considered as contributing to the overall thickness of soils
required for frost protection.
The final depth of footing embedment may be controlled by the calculated
scour depth and be deeper than the depth required for frost protection. Refer
to Section 2.3.11 Scour for information regarding scour depth.
5.2.2 Seal Cofferdams
Seal cofferdams are used when a substructure unit must be constructed with
its foundation more than 4 feet below the water table, to counteract the
buoyant forces produced during pumping of the cofferdam. Once the
cofferdam is constructed, the seal is placed under water and water is then
pumped out of the cofferdam. This provides a dry platform for construction of
the spread footing, or in the case of a pile foundation, the distribution slab.
When a seal is needed, the top of footing or distribution slab is located
approximately at streambed, and the depth of seal is calculated based upon
the buoyancy of the concrete under the expected water surface during
construction. The following formula can be used:
145 y 62.4 z

where:
145 lb/ft3 =
62.4 lb/ft3 =
y=
z=

March 2014

unit weight of concrete


unit weight of water
the depth of seal from top of seal to bottom of seal
the depth of water from water surface to bottom of seal

5-6

CHAPTER 5 - SUBSTRUCTURES

The depth of water in the above formula should be based on an appropriate


flood event, but no less than Q10. The depth of water at tidal locations should
be selected on a case-by-case basis, but no less than MHW. A note should
be included on seal cofferdam sheets specifying the water elevation assumed
in the design and specifying adjusting the seal depth should the water
elevation at the time of construction be higher. To prevent seal buoyancy
during a high water event after construction is complete, the Designer may
specify vent holes at the design height of water, on a case-by-case basis.
Anchorage of the footing or distribution slab to the seal is required. For pilesupported foundations, this can be accomplished by extending the piles into
the distribution slab. For seals founded on bedrock, dowels should be drilled
and grouted into the seal after dewatering and prior to placement of the
footing.
When sheet piling is used for a seal cofferdam, the minimum dimensions for
the seal should be shown on the design drawings. These dimensions and
details should be noted on the plans in conjunction with the appropriate notes
in Appendix D Standard Notes Seal Cofferdams.
5.2.3 Cofferdams
Cofferdams are retaining structures with the retained material being water and
soil. A separate cofferdam must be specified for the construction of each
substructure unit (abutment or pier) that cannot be constructed completely in
the dry. When water cannot be controlled so that footing concrete can be
placed in the dry, a concrete seal must be placed below the elevation of the
footing. Refer to Section 5.2.2 Seal Cofferdams.
Cofferdam design is the responsibility of the Contractor, and construction
requirements are found in Standard Specification Section 511 Cofferdams.
Unless otherwise provided or approved, cofferdams are removed after the
completion of the substructure, with care being taken not to disturb or
otherwise damage the finished work.
Cofferdams should not be specified for substructure units that are constructed
on dry land, such as on overpass structures. For large braced excavations a
Special Provision should be included in the PS&E package to pay for braced
excavations under the appropriate cofferdam item. Any temporary retaining
structures that are required to support small structural excavations should be
considered incidental to the appropriate structural excavation or substructure
pay items.
Cofferdam requirements for culverts and other buried structures are found in
Section 8.1.2 Construction Practices.

March 2014

5-7

CHAPTER 5 - SUBSTRUCTURES

5.2.4 Concrete Joints


Concrete joints in a vertical plane are used in concrete construction to
accommodate changes in the volume of concrete caused by such factors as
drying shrinkage, creep, and the application of load. When concrete is
restrained by internal or external forces, the stresses caused by concrete
movement would be relieved by the formation of significant cracks, if joints
were not provided. Construction joints are used to facilitate the sequence of
construction, and are typically located in a horizontal plane for abutments,
piers, and walls.
There are three types of joints commonly used in concrete construction. A
concrete key is generally used with each joint for shear transfer, as shown in
Standard Detail 502 (01). The Structural Designer should specify the proper
concrete joint, depending upon its intended use.

Contraction joints are normally used every 30 feet along a wall to


control the location of cracks. Without these joints, the concrete
would form cracks at unpredictable intervals. Reinforcing steel is
normally not carried through the joint, except in rigid frame structures,
where moment must be transferred from wall to slab.

Expansion joints are used to prevent compression forces from


abutting concrete from crushing or displacing the adjacent structure.
It is good practice to locate expansion joints where expansion forces
change direction, such as at wingwall turns. In retaining walls and
abutment/wingwall systems, expansion joints should be spaced no
more than 90 feet apart. Reinforcing steel is not carried through the
joint.

Construction joints are used between concrete placements when the


sequence of construction requires more than one placement. The
surface between placements becomes a construction joint. These
joints may be designed to coincide with contraction or expansion
joints. If not functioning as a contraction or expansion joint,
reinforcing steel is normally carried through the joint.

A horizontal construction joint in the abutment backwall should be


shown on the plans to facilitate installation of the superstructure
expansion device. This should normally be located at a minimum
vertical distance of 1-3 from the roadway surface, except for
modular expansion devices, which must conform to the
manufacturers recommendations (refer to Section 4.8.5 Modular
Joints). Bent #5 bars at 1-6 maximum spacing should be used in
the top of the backwall. Welding to reinforcing steel is allowed in this
area so that the Contractor can utilize the reinforcing steel to support
the expansion device.

March 2014

5-8

CHAPTER 5 - SUBSTRUCTURES

5.2.5 Seismic Considerations


Seismic analysis of bridges and foundations shall be performed in accordance
with the LRFD Specifications or the AASHTO Guide Specifications for LRFD
Seismic Design (herein referred to as the Guide Specification).
Seismic analysis is not required for the following:

Single-span bridges, regardless of Seismic Zone or Seismic Design


Category (SDC)

Any bridge in Seismic Zone 1 or SDC A, with the exceptions


described below.

For all bridges, including those for which seismic analysis is not required,
superstructure connections and bridge seat dimensions should be satisfied per
LRFD 3.10.9 and 4.7.4.4, respectively.
For critical or essential bridges, including those in Seismic Zone 1 or SDC A,
the Department may specify a higher Seismic Zone or SDC than that specified
by the LRFD Specifications and the Guide Specification or specify appropriate
seismic provisions. Critical and essential bridges are not specifically
classified in this Bridge Design Guide, but will be designated as such by the
Department at its discretion.
In general, bridges that may be classified by the Department as critical or
essential are as follows:

Bridges that are required to be open to all traffic once inspected after
the design earthquake and usable by emergency vehicles and for
security, defense, economic or secondary life safety purposes
immediately after the design earthquake.

Bridges that should be open to emergency vehicles and/or for


security, defense or economic purposes after the design earthquake
and open to all traffic within days after that event.

Bridges that are formally designated as critical for a defined local


emergency plan.

For non-conventional bridges, including cable-stayed and suspension bridges,


truss bridges, arch type bridges and movable bridges the Department will
specify and approve appropriate seismic design provisions.
It is estimated that most bridge sites in Maine will be classified as Seismic
Zone 1 or SPC A. The exception are bridges in the extreme northwest portion
where the subsurface conditions might be classified as Site Class B, C or D,
and bridge sites everywhere where the subsurface conditions are Site Class
March 2014

5-9

CHAPTER 5 - SUBSTRUCTURES

E, except those in downeast coastal Maine. It is estimated these bridge sites


will be classified as Seismic Zone 2 or SDC B.
For bridges requiring seismic analysis, the effect of earthquake loading on the
foundations shall be investigated using the extreme event limit state in LRFD
Table 3.4.1-1 with resistance factors, , of 1.0 and an appropriate seismic
analysis method as described in LRFD 4.7.4.3 and LRFD 3.10.9.2 through
3.10.9.4. The foundation design should consider the effect of wall inertia and
amplification of active earth pressure by earthquake determined by the
Mononobe-Okabe method. The Mononobe-Okabe method for determining
equivalent static fluid pressure for seismic loads on walls is presented in LRFD
11.6.5 and Appendix A11. LRFD Appendix A10 gives additional guidance
regarding seismic analysis and design of foundations.
For foundations on soil and rock, the location of the resultant of the reaction
forces due to earthquake loading should be within the middle two-thirds (2/3)
of the footing base for EQ = 0.0 and within the middle eight-tenths (8/10) of the
footing base for EQ = 1.0. For in between values of EQ, the restriction for the
location of the resultant is obtained by linear interpolation of the preceding
values of EQ.
For overall stability of a retaining wall when earthquake loading is included, a
resistance factor, , of 0.90 should be used. For bearing resistance, a
resistance factor, , of 0.80 should be used for gravity and semigravity walls
and 0.90 for MSE walls.
Where the backfill or foundation soils are saturated, consideration should be
given to address the possibility of soil liquefaction and lateral spreading.
Liquefaction design guidance is provided in LRFD 10.5.4.2, 11.5.4.2 and
Appendix A10.
5.3

Spread Footings

Spread footings should be designed and proportioned for the strength, service,
and extreme event limit states such that the factored resistance is not less that
the effects of the factored loads specified in LRFD Article 3.
Selection of foundation type is based on an assessment of the magnitude and
direction of loading, depth to suitable bearing materials, flood history, potential for
liquefaction, undermining, scour or wave action, frost depth, and ease and cost of
construction.
5.3.1 Service Limit States
Spread footings at the service limit state shall be investigated for:

Settlement

March 2014

5-10

CHAPTER 5 - SUBSTRUCTURES

Horizontal movement

Rotation

Overall stability of slope with the footing

Scour at the design flood, specified in LFRD 2.6.4.4.2 and 3.7.5

Settlement shall be investigated for the Service I Load Combination and


rotations and horizontal movements shall be investigated at all applicable
service limit states.
The tolerable level of ultimate settlement, differential settlement, rotation and
horizontal movement shall be controlled by superstructure tolerance,
rideability, span length, road classification, long-term maintainability and
economy.
Bearing resistance estimated using presumptive allowable bearing resistances
shall only be applied to address service limit state load combinations or for
preliminary sizing of footings.
Service limit state analyses shall use unfactored loads. Resistance factors for
the service limit state shall be taken as 1.0. The exception is the investigation
of the overall slope stability of a retaining wall or an earth slope supporting a
retaining wall footing or an abutment footing. In those instances, the earth
slopes should be investigated at the Service I Load Combination, with a
resistance factor,of 0.65.
5.3.2 Strength Limit States
The design of spread footings at the strength limit states shall consider:

Factored bearing resistance

Eccentricity or loss of contact

Sliding

Loss of lateral and vertical support due to scour at the design flood
event; the design flood is defined as the more severe of the 100-year
even or an overtopping flood of lesser recurrence interval.

Factored structural resistance

Resistance factors for the bearing resistance of spread footings at the strength
limit state are provided in Section 5.3.5.3. Resistance factors for sliding are
provided in Section 5.3.8.

March 2014

5-11

CHAPTER 5 - SUBSTRUCTURES

A modified Strength Limit State analysis should be performed that includes in


the ice pressures specified in Section 3.9 Ice Loads, with the appropriate
strength limit state resistance factors. That Strength Limit State that results in
the extreme force and moment effects should be selected.
5.3.3 Extreme Event Limit States
Spread footings should be designed for extreme events such as seismic
loads, liquefaction, check flood for scour, vessel impact, vehicle or railway
collision, and ice.
The ice pressures for the Extreme Event II Limit State should be unfactored
and applied at Q1.1 and Q50 elevations as defined in Section 3.9 Ice Loads
but with the ice thickness increased by 1 foot.
Resistance factors for extreme event limit states shall be taken as 1.0.
For the extreme event limit state, the Designer should consider scour due to
the check flood event and should determine that there is adequate foundation
resistance to support all applicable unfactored loads with a resistance factor of
1.0. Flood event loads should include debris loads, where applicable.
Extreme limit state design checks for spread footings shall include checks of:

Bearing resistance

Eccentricity

Sliding

Overall stability

5.3.4 Footing Depth


Footings should be embedded a sufficient depth to provide adequate bearing
materials and protection against frost action, erosion and scour.
5.3.4.1

Bearing Materials

A footing should ideally be founded on a single material type throughout


its bearing length. If a combination of materials is present underlying the
footing (i.e., bedrock and granular material) the granular material should
be removed to the bedrock surface and replaced with concrete fill. In
special situations where constructing a footing on dissimilar materials
cannot be avoided, see the Geotechnical Designer.

March 2014

5-12

CHAPTER 5 - SUBSTRUCTURES

Footings should be founded on firm soils or bedrock. Any organic, loose,


or otherwise unsuitable material encountered at the footing elevation
should be removed to the full depth and replaced with compacted granular
fill or concrete fill to the bottom of footing elevation. If concrete fill is used
under a foundation, the pay limits should be shown as a vertical plane and
should be designated as "Pay Limit for Structural Excavation and
Concrete Fill". The distance outside the footing for the concrete fill pay
limit should be determined for each individual case and must be shown on
the design drawings. Foundation bearing conditions should be approved
in the field by the Construction Resident or Geotechnical Designer.
5.3.4.2

Footings on Bedrock

For footings supported on bedrock the surface will be cleaned of all


weathered bedrock, fractured material, loose soil, and/or ponded water
prior to placement of the footing concrete. Smooth bedrock should be
roughened or serrated prior to placing concrete to enhance sliding
stability. The foundation bearing areas should be approximately level.
Bedrock slopes that exceed 4H:1V should be step-serrated or suitably
benched to create level steps or a completely level subgrade. For
bedrock slopes between 4H:1V and 6H:1V consider dowels into bedrock
to control sliding potential.
5.3.4.3

Frost Protection

Footings will be placed below the depth of frost penetration as discussed


in Section 5.2.1 Frost. Riprap is not to be considered as contributing to
the overall thickness of soils required for frost protection.
5.3.4.4

Scour Protection

Spread footings on soil or erodible rock at stream crossings should be


founded at a depth at least 2 feet below scour depth of scour determined
for the check flood for scour. Spread footings supported on soil within the
stream channel shall be located a minimum of 6 feet below the thalweg of
the waterway. Refer to Section 2.3.11 Scour for information regarding
scour depth.
5.3.5 Bearing Resistance
5.3.5.1

General

Spread footings for abutments and retaining walls are to be proportioned


to ensure stability against bearing capacity failure. Safety against deep
seated foundation failure shall also be investigated per LRFD 10.6.2.3.

March 2014

5-13

CHAPTER 5 - SUBSTRUCTURES

Bearing resistance should be investigated at the strength limit state.


LRFD Article 11.6.3.2 and Figures 11.6.3.2-1 and -2 provide examples for
calculating the vertical bearing stress. In general, load factors selected
should produce the total extreme force effect. Specific guidance for
selection of load factors for bearing resistance is provided in LRFD Figure
C11.5.6-1. Where there is a live load surcharge, the factored surcharge
force is included over the backfill immediately above the wall base or
footing.
Spread footings should be designed such that the factored design stress
does not exceed the factored bearing resistance of the soil or rock. The
nominal bearing resistance of footings on soil may be estimated using the
Munfakh procedure outlined in LRFD Article 10.6.3.1.2. The use of
Terzaghi, Meyerhof, or Vesic methods for estimating the nominal bearing
resistance is also acceptable. Consideration of shape factors, inclined
loads, ground surface slope, and eccentric loading should be included in
the calculation, if applicable. A resistance factor shall be applied to the
calculated nominal resistance. Structures should be designed such that
the maximum factored pressure on the soil or rock under footings does not
exceed the factored bearing resistance provided by the Geotechnical
Designer.
The bearing resistance at the service limit state will be settlement
controlled (typically 1 inch). Presumptive bearing resistance charts based
on soil or rock type may be used to determine the service limit state
bearing resistance.
For spread footings on bedrock, the design of the footing is typically
controlled by overall stability, i.e., failure along discontinuities in the rock
mass or eccentricity. Therefore, the Designer should verify overall
stability by sizing the footing based on eccentricity at the strength limit
state and then checking the nominal bearing resistance at the service and
strength limit states.
5.3.5.2

Bearing Stress Distribution

The distribution of soil pressure should be consistent with the foundation


material, whether it is soil or bedrock. When proportioning footing
dimensions to meet settlement and bearing resistance requirements, the
distribution of bearing stress on the effective footing area shall be
assumed to be:

March 2014

Uniform for footings on soils

Triangular or trapezoidal for footings on rock

5-14

CHAPTER 5 - SUBSTRUCTURES

For structural design of footings, a triangular or trapezoidal stress


distribution based on factored loads should be used regardless if the
footing bears on soil or rock.
When loads are eccentric, the bearing stress is distributed to the effective
footing area, L x B, where the reduced dimensions are taken as:

B = B - 2eB

L = L - 2eL

where eB and eL are the eccentricities relative to a point at the center of


the footing, parallel to the B and L dimensions, respectively.
5.3.5.3

Bearing Resistance Factors

The resistance factors for bearing resistance are provided in Table 5-2.

March 2014

5-15

CHAPTER 5 - SUBSTRUCTURES

Table 5-2 Bearing Resistance Factors


Method/Soil/Condition
Theoretical method (Munfakh et al. 2001) in
clay
Theoretical method (Munfakh et al. 2001) in
sand using SPT
Semi-empirical methods (Meyerhof, 1957,
Terzaghi, Vesic) all soils
Footings on rock
Plate Load Test

Bearing
Resistance
Factor, b
0.50
0.45
0.45
0.45
0.50

5.3.6 Settlement
The design of spread footings is frequently controlled by settlement at the
service limit state. It is advantageous to proportion spread footings at the
service limit state and check for adequate design at the strength and extreme
limit states.
Total and differential settlement should be evaluated. The total settlement
includes elastic settlement, primary consolidation, and secondary
compression. Elastic settlement results from the compression of the material
supporting the foundation or from reduction in pore space in nonsaturated
soils. Consolidation settlement occurs when saturated, fine-grained soils
experience an increase in stress. Some soils, after experiencing primary
consolidation settlement, continue to strain after excess pore-water pressures
are dissipated. This process is termed secondary compression, or creep.
Immediate or elastic settlement should be determined using the Service I Load
Combination, specified as unfactored dead load, plus the unfactored
component of live loads assumed to extend to the footing level. Timedependent settlements, i.e., primary consolidation and secondary compression
settlement may be determined using the unfactored dead load only. Other
factors that can affect settlement, such as embankment loading, lateral and/or
eccentric loading, and dynamic or earthquake loads should also be
considered, where applicable.
Differential settlement occurs when one load-bearing member of a structure
experiences total settlement of a different magnitude than an adjacent loadbearing member. Transportation structures, especially bridges, are not
exceptionally tolerant of differential settlements. Deformation limitations will
form the upper bound of allowable differential settlements used to design
shallow foundations.
March 2014

5-16

CHAPTER 5 - SUBSTRUCTURES

5.3.6.1

Tolerable Settlement

Foundation settlement criteria should be consistent with:

The type of structure

The function of the structure

Anticipated service life

Consequences of unacceptable movement on structure


performance

Long-term maintainability

Tolerable movements are frequently described in terms of angular


distortion between members. Angular distortion ('/) between adjacent
foundations should be limited to 0.008 radians for simple span bridges and
0.004 radians for continuous span bridges, where ' is the differential
settlement and is the span length. Angular distortion limits may deviate
on a project by project basis, depending on:

The cost of mitigating settlement through larger foundations,


realignment, lightweight fills or surcharge

Rideability

Aesthetics

Safety

Tolerance of the superstructure to lateral movement will depend on the


bridge seat or joint widths, bearing type and structure type.
5.3.6.2

Settlement Analyses

Settlement may be estimated using procedures described in LRFD


10.6.2.4 or other generally accepted methods. The soil parameters used
shall be based on the results of laboratory or insitu testing, or both. Total
and differential settlement should be evaluated.
Settlement of spread footings on sand can be predicted using calculation
methods by Hough, Peck-Bazaraa, DAppolonia, or Schmertmann, as
applicable.

March 2014

5-17

CHAPTER 5 - SUBSTRUCTURES

5.3.7 Overall Stability


The overall global stability of spread footings on or near an earth slope should
be investigated using Service I Load Combination and an appropriate
resistance factor. Where a slope supports or contains a structural element,
such as a spread footing supporting a wall or abutment, the resistance factor,
, shall be taken as of 0.65
For foundations on spread footings constructed along rivers and streams,
scour of foundation materials is evaluated as specified in LRFD 2.6.4.4.2.
Extreme limit state design should check that the nominal resistance of the
footing and slope remaining after the scour due to the check flood for scour
can support the unfactored strength limit state loads with a resistance factor,
, of 1.0
The overall stability of retaining wall spread footings on or near a slope should
be evaluated using limiting equilibrium methods of analysis, which employ the
Modified Bishop, simplified Janbu, Spencer, or other generally accepted
methods of slope stability analysis.
5.3.8 Sliding
Failure by sliding should be investigated for all spread footings bearing on soil
or bedrock. Passive earth pressure exerted by fill in front of the footing should
be neglected in consideration that the soil may be removed as the result of
scour or during future construction, and in consideration that soils in front of
the footing will be subject to freeze-thaw weakening over time. If passive
pressure is included as part of shear resistance to sliding, consideration
should be made to possible removal of the soil in front of the foundation in the
future. If passive resistance is included in the resistance, its magnitude is
commonly taken as 50% of the maximum passive pressure resistance
computed using Rankine Passive resistance. This is the basis of a resistance
factor for passive resistance of ep of 0.50.
The factored resistance against failure by sliding is taken as:
Rr = Rn = sRf + epRep
where:
Rn = nominal sliding resistance
s = resistance factor for shear resistance between soil and
foundation specified in Table 5-3.
Rf = nominal sliding resistance between soil and foundation
ep = resistance factor for passive resistance = 0.50
Rep = nominal passive resistance of the soil available throughout
the design life of the structure.

March 2014

5-18

CHAPTER 5 - SUBSTRUCTURES

Table 5-3 Resistance Factors for Sliding of Spread Footings at


the Strength Limit State
Soil/Condition
Precast concrete on sand
Cast-in-place concrete on sand
Cast-in-place or precast concrete on clay
Soil on soil
Cast-in-place concrete on rock (based on
reliability theory analysis of footings on sand)
Cast-in-place concrete on rock (calibrated to ASD
Factor of Safety of 1.5)

Sliding
Resistance
Factor, s
0.90
0.80
0.85
0.90
0.80
0.90

Spread footings should be designed such that the factored resistance to


sliding, Rf, is greater than the factored force effects due to the horizontal
components of loads. Load factors selected should produce the extreme force
effect. The live load surcharge is not included over the heel. Specific
guidance for selection of load factors for sliding are provided in LRFD Figure
C11.5.6-2.
The nominal sliding resistance between footings and cohesionless soils is
taken as:
Rf = V x tan
where:
tan = tan for cast-in-place footings on soil
tan = 0.80 tan for precast footings on soil
V = total vertical force
The coefficient of friction, tan , for sliding should be as shown in Table 3-3 for
the soil type under the footing and LRFD Table 3.11.5.3-1.
The nominal sliding resistance between footings and silt and/or clay soils
should be taken to be the lesser of: (1) the undrained shear strength of the
silt/clay, or, (2) one-half of the normal stress on soil when the footing is
founded on at least 6 inches of compacted granular fill on silt/clay.
For footings on bedrock, the Geotechnical Designer will provide a coefficient of
friction for sliding. If smooth bedrock is present at the bearing elevation or if
the coefficient of sliding is insufficient to resist lateral forces, the bedrock
should be doweled to improve stability. When a footing is doweled into rock,
the dowels should be #9 reinforcing bars or larger and be embedded into the
footings and bedrock by depths determined by the Designer. The spacing of

March 2014

5-19

CHAPTER 5 - SUBSTRUCTURES

the dowels should be no greater than 3 feet between rows and no less than
two rows. If sloping bedrock is present (steeper than 4H:1V) at the bearing
elevation, the bedrock should be benched to create level steps or doweled to
improve stability.
5.3.9 Eccentricity
Load factors for eccentricity selected should produce the extreme force effect.
The live load surcharge is not included over the heel of the footing. Specific
guidance for selection of load factors for eccentricity are provided in LRFD
Figure C11.5.5-2. The location of the resultant of the reaction forces shall be:

within the middle two-thirds (2/3) of the footing width or length, B or L,


for footings on soils, or

within the middle nine-tens (9/10) of the footing width or length, B or


L, for footings on rock.

5.3.10 Ground Water Condition


Footing excavations below the ground water table, particularly in granular soils
having relatively high permeability, should be made such that the hydraulic
gradient in the excavation bottom is not increased to a magnitude that would
cause the foundation soils to loosen or soften due to upward flow of water.
Dewatering or cutoff measures to control seepage should be used where
necessary. Footing design should be calculated using the highest anticipated
ground water level at the footing location.
5.3.11 Drainage Considerations
Adequate drainage of materials behind structures is of great importance and
should be provided as described in Section 5.4.1.9 Drainage.
5.4

Abutments

5.4.1 Conventional Abutments


5.4.1.1

General Design Requirements

Abutment and wingwall design should include evaluation of settlement,


lateral displacement, overall stability of the earth slope with the foundation
unit, bearing capacity, sliding, loss of contact with foundation soils,
eccentricity (overturning), pile capacity (if applicable) and structural
capacity. Abutments should be designed for extreme events such as
vessel collisions, vehicle collisions, and seismic activities, along with

March 2014

5-20

CHAPTER 5 - SUBSTRUCTURES

changed conditions such as scour, as applicable. The design of


abutments and walls should satisfy service, strength, and extreme limit
state requirements.
5.4.1.2

Loads Combinations and Load Factors

Structural analyses and geotechnical evaluation of abutments should be


performed in accordance with the AASHTO LRFD. Abutments should be
designed and proportioned to resist all applicable load combinations
specified in LRFD Articles 3.4.1 and 11.5.5 and as outlined in Chapter 3
Loads.
Abutments should be evaluated for each of the applicable limit states:

March 2014

Strength I-construction. Strength Limit State I with the


exception that bridge superstructure DC and DW, and vehicular
live loads, LL, are neglected. Load factors for the dead load of
other components shall not be less than 1.25. Live load
surcharge is included to account for construction equipment live
loading during structure erection and a construction load factor
of not less than 1.5 should be assumed. The Strength Iconstruction analysis should investigate any anticipated
construction loadings, such as looking at the abutment partially
backfilled without the superstructure in place.

Strength I-a: Strength Limit State I, which models the basic load
combination related to normal vehicle use of the bridge without
wind, dead load plus earth pressure, finished grade, including
the vertical component of the superstructure, approach slab, live
load effects of traffic on the approach (LS) the vertical
component of the live load from superstructure. Minimum
vertical permanent load factors and maximum horizontal load
factors are selected to produce extreme force effects for
abutment sliding and eccentricity, and structural design of the
abutment stem.

Strength 1-b: Strength Limit State 1 as described above, except


maximum vertical permanent load factors, including earth loads,
are selected to produce an extreme force effect for bearing
capacity analyses.

Strength III: Load combination relating to the bridge exposed to


high wind velocity (100 mph) without live loads. Minimum and
maximum load factors should be selected for permanent loads
to investigate the most extreme force or moment effect.

5-21

CHAPTER 5 - SUBSTRUCTURES

Strength IV: Load combination relating to very high dead load to


live load force effect ratios exceeding about 7.0 Strength IV will
likely govern for bearing failure on long span bridges. It also
will likely govern for structural design of the footing. Minimum
and maximum load factors should be selected for Permanent
Loads to investigate the most extreme force or moment effect.

Strength V: Load combination relating to the bride exposed to


wind velocity of 55 mph with live loads. Minimum and
maximum load factors should be selected for permanent loads
to investigate the most extreme force or moment effect.

Service I: Service Limit State I Load combination relating to


normal operational use of the bridge with a 55 mph wind and all
loads taken at their unfactored values.

For the load combinations with all dead loads applied, with or without the
superstructure live load, distribute the superstructure loads over the length
of the abutment between the fascia lines of the superstructure.
Where abutments are to be designed to resist earthquake forces,
collisions by roadway or rail vehicles, or vessel collision, the structures
should be evaluated for the following additional limit states:

Extreme Event I Load combination including earthquake


forces

Extreme Event II Load combination relating to collision by


vehicles or vessels.

Certain permanent loads, including earth loads, should be factored using


the load factors p. Permanent load factors should be selected to produce
the total extreme factored force effect. Typical load factors, load
combinations and the analyses for which they will govern, are provided in
Table 5-4.

March 2014

5-22

CHAPTER 5 - SUBSTRUCTURES

Table 5-4 Typical Load Groups and Load Factors (i) for
Abutments on Spread Footings
Controlling
Load
Group

DC

EV

Ls

EH
(active
or
passive)

Strength I-a

0.90

1.0

1.75

1.5

Strength I-b

1.25

1.35

1.75

1.5

LL

Analysis
Governed

- Sliding
- Eccentricity
1.75
(overturning)
- Structural design
of wall stem
- Bearing Capacity
1.75

Strength IV

1.50

1.35

--

1.5

--

Service I

1.0

1.0

1.0

1.0

1.0

- Bearing capacity
- Structural design
of the footing
- Settlement
- Lateral movement
- Angular distortion

Longitudinal forces for abutment design should include any live load
longitudinal forces developed through bearings such as braking forces, or
others as specified in LRFD Article 3.0, unless limited by friction capacity.
5.4.1.3

General

The Designer should estimate the load combinations which could be


imposed on the abutment or wall and estimate the nominal resistance of
the structural component or ground. Abutment components shall satisfy
the following equation for each limit state:
i i Qi Rn = Rf
where:
i = Factors to account for ductility, redundancy and operational
importance
i = Load factor (dim)
Qi = Load or stress
= Resistance factor (dim)
Rn = Nominal resistance
Rf = Factored resistance

March 2014

5-23

CHAPTER 5 - SUBSTRUCTURES

5.4.1.4

Strength Limit State Evaluations

The above equation should be used to evaluate abutments at the strength


limit states for:

Bearing resistance failure

Lateral sliding

Excessive loss of base contact (eccentricity)

Pile failure

Structural failure

The factored resistance, Rf, calculated for each mode of failure, is to be


calculated using the appropriate resistance factors for bearing resistance,
sliding, eccentricity, axial pile resistance and structural resistance.
The Designer should consider the consequences of changes in abutment
foundation conditions at the strength limit state resulting from scour due to
the design flood event using appropriate resistance factors.
5.4.1.5

Service Limit State Evaluations

Abutments should be investigated at the service limit state using the load
and resistance equation in Section 5.4.1.3 for:

Settlement

Lateral displacement

Overall slope stability

Overall stability at the design flood

A resistance factor, , of 1.0 is used to assess abutment design at the


service limit state. Overall stability of abutments on or near earth slopes
should be investigated using resistance factors in Section 5.3.7 Overall
Stability.
Tolerable vertical and lateral displacement criteria for abutment shall be
developed based on the function and type of wall, anticipated service life,
and consequences of unacceptable movements of the wall and effect on
nearby structures. To control bridge superstructure damage, a limiting
horizontal movement of abutments less than 1.5 inch is recommended.
Utilities may not be able to accommodate very large movements, in which
case a project-specific limiting movement should be developed.

March 2014

5-24

CHAPTER 5 - SUBSTRUCTURES

5.4.1.6

Extreme Limit State Evaluations

Extreme limit state design checks for abutments should include:

Bearing resistance

Eccentricity

Sliding

Overall stability

A resistance factor, , of 1.0 is used in the load and resistance equation in


Section 5.4.1.3 to assess abutment design at the extreme limit state.
The extreme event limit state design should check that the nominal
abutment foundation resistance after scour due to the check flood event
can support all applicable unfactored loads with a resistance factor of 1.0.
For abutments on spread footings, refer to 5.3.4.4. For pile-supported
abutments, refer to 5.4.1.12.
5.4.1.7

Load Considerations

A. Earth Loads
For abutment and wingwall designs, use the appropriate soil weight
shown for Soil Type 4 (Table 3-3) for soil properties for backfill material.
Abutments and retaining walls should be designed as unrestrained and
free to rotate at the top in an active state of earth pressure. An active
earth pressure coefficient, Ka, should be calculated using Rankine
Theory for long-heeled cantilever abutments and wingwalls, and
Coulomb Theory for short heeled cantilever abutments and gravity
shaped walls. Refer to Section 3.6.5.1 Coulomb Theory. Soil Type 4
properties are consistent with materials typically used for backfill behind
abutments and retaining walls. For unconventional backfills, i.e., tire
shreds, light weight fills, etc., consult the Geotechnical Designer or
Report.
B. Unit Weight of Concrete
A unit weight of 150 lb/ft3 should be used for design purposes.
C. Live Load Surcharge Loads
Abutments without approach slabs should be designed with a live load
surcharge when computing horizontal earth pressure. This additional
lateral pressure on walls is approximated by a uniform horizontal earth
pressure due to an equivalent height of soil, Heq. Refer to Section 3.6.8
March 2014

5-25

CHAPTER 5 - SUBSTRUCTURES

Surcharge Loads for guidance in computing this additional lateral


surcharge pressure.
Wingwalls and retaining walls should also be designed for surcharge
loads in accordance with Section 3.6.8.
In the case a structural approach slab is specified, reduction, but not
elimination, of the surcharge loads is permitted per LRFD 3.11.6.2.
D. Lateral Loads
Load conditions should include any additional lateral pressures on the
walls. These loads may include but are not limited to impact loads
transmitted to the retaining walls from distribution slabs supporting crash
barriers.
E. Collision Forces
Unless the department determines that site conditions indicate
otherwise, abutments within a distance of 30 feet to the edge of a
roadway or within 50 feet to the centerline of railway track shall be
investigated for collision. Collision loads and crashworthy barrier design
criteria for abutments are identical to those provided for Piers in Section
5.5.1.10 Pier Protection.
5.4.1.8

Backfill

Abutment walls and footings should be backfilled with granular borrow for
underwater backfill. Extend underwater granular backfill for a horizontal
distance of at least 10 feet from the back face of the abutment wall and 1
foot behind the back face of the footings.
5.4.1.9

Drainage

The Designer should study total drainage design. Adequate drainage of


fill behind structures is important to increase the longevity of retaining
structures. Water should not drain into the underside of slope protection.
Drainage should be provided as follows:

March 2014

Where possible, french drains should be used at the back face


of walls with 4 inch diameter drain pipes (weep holes) at
nominal 10 foot maximum spacing through the walls . Refer to
Standard Specification Section 512 French Drains.

Underdrains or other means may be used where necessary to


provide adequate drainage.

5-26

CHAPTER 5 - SUBSTRUCTURES

5.4.1.10 Reinforcement and Structural Design


The structural design of abutments should comply with the requirements
of AASHTO LRFD. Earth loads for structural design should be calculated
per Section 3.4, Earth Loads, and an appropriate load factor applied.
Concrete cover for footing reinforcement should be as specified by
AASHTO LRFD, except that for "non-designed" footings, such as for stub
abutments, 6 inches of cover should be used.
At the back corners of gravity abutments and wingwalls, horizontal rebar
should be placed, #6 bars at 12 inches on center, with lengths of 8 feet
and with 6 inches of cover. Also, four #6 bars, 8 feet long, should be
placed at 6 inches below bridge seat elevation at the front corners.
5.4.1.11 Abutments on Spread Footings
A. General
Refer to Section 5.3 Spread Footings for guidance on the design of
spread footings.
The general design process for spread footing design should follow the
steps below:
1. Determine the nominal and factored footing resistances at the
service, strength and extreme limit states assuming footing
dimensions and depth (consult Geotechnical Design Report)
2. Determine the loads applied to the footing, including lateral earth
pressure loads for the abutment
3. Initially size and design the footing at the service limit state
4. Check the bearing pressure of the footing at the strength limit
state
5. Check the eccentricity of the footing at the strength limit state
6. Check the sliding resistance of the footing at the strength limit
state
7. Check the bearing pressure and eccentricity and sliding
resistance of the footing at the extreme limit state
8. Check the footing bearing resistance at all limit states and overall
stability in light any refined/new footing dimensions, depth and
loads provided by the Designer.
March 2014

5-27

CHAPTER 5 - SUBSTRUCTURES

9. Reassess steps 4 thru 7 based on the revised nominal and


factored footing bearing resistance calculated
B. Spread Footings on Bedrock
Refer to Section 5.3.4.2 for guidance on the design of spread footings on
bedrock.
C. Vertical and Horizontal Displacement
Vertical and horizontal movement criteria for abutments should be
developed consistent with the function and type of structure,
consequences of unacceptable movements on structure performance
and the cost of mitigating movements and/or rotations by larger
foundations. Angular distortions and settlements should be designed
per Section 5.3.6 Settlement.
D. Global Stability
Global stability of slopes with abutments or walls should be considered
part of the design of the wall or abutment. Evaluation of the global
stability of an abutment is important when the abutment is located close
to or on an inclined slope, or close to an embankment, excavation, or
retaining wall.
The evaluation of the overall stability of earth or rock slopes with walls
and abutments shall be investigated at the Service I Load Combination
and a resistance factor, , of 0.65. Refer to 5.3.7 Overall Stability for
additional guidance.
E. Bearing Stress
Maximum bearing stress under footings at the strength limit load
combination should be determined per Section 5.3.5 Bearing
Resistance. Structures should be designed such that the calculated
factored bearing stress under footings does not exceed the factored soil
or rock bearing resistance in accordance with recommendations of the
Geotechnical Designer. This requirement is expressed below:
Rn = Rf
where:
= factored vertical stress (ksf)
= bearing resistance factor (dim)
Rn = nominal bearing resistance (ksf)
Rf = factored bearing resistance (ksf)

March 2014

5-28

CHAPTER 5 - SUBSTRUCTURES

The weight of the earth in front of a wall should be considered in


computing maximum bearing pressure. When loads are eccentric, the
effective footing dimension should be used for the overall dimension in
the equation for bearing resistance. Refer to Procedure 5-1 and
Procedure 5-2 for how to calculate applied bearing stress.

March 2014

5-29

CHAPTER 5 - SUBSTRUCTURES

Procedure 5-1 Bearing Stress on Soil


For Wall or Conventional Abutment
CL Base (Footing)

LS

DLv, LLv

XDL, x LL

W1

PLS

V1

Xv1

Ph

Xw1

H/3

W2

B/2
B 2e
B

Step 1. Calculate eccentricity, ec, about point C, where:


Mo = sum of moments of factored overturning forces acting about point C:

M o Ph

H
H
W1 x w1 PLS ,h DLV x DL LLv x LL
3
2

Mr = sum of moments of factored resisting forces acting about Point C:

M r V1 xV 1 LS L x V 1
V = sum of factored vertical forces acting on the footing and wall:

March 2014

5-30

CHAPTER 5 - SUBSTRUCTURES

V V1 W1 W2 DLv LLv LS L
and,

ec

ec

PH

Mo Mr
V

H
H
V1 xV 1 LS L xV 1 W1 xW 1 PLS , h DLv x DL LLV x LL
3
2
V1 W1 W2 DLv LLv LS L

Step 2. The factored vertical stress should be calculated assuming a uniformly


distributed pressure over an effective base area shown in the Figure above. The vertical
stress should be calculated as follows:

V
B 2e c

Note that B-2ec is considered to be the effective footing width.


Step 3: Compare v which already has the load factors included, to the factored bearing
resistance of the soil, provided in the Geotechnical Report. The maximum factored
stress should be less than the factored bearing resistance.

bc Rn R f

where:
v = factored vertical stress (ksf)
bc = bearing resistance factor (dim)
Rn = nominal bearing resistance (ksf)
Rf = factored bearing resistance (ksf)
Note: The case shown for this procedure is the construction load with full backfill and live
load surcharge on the approach, and superstructure dead load. For other load
combinations, the appropriate loads must be included in the analysis.

March 2014

5-31

CHAPTER 5 - SUBSTRUCTURES

Procedure 5-2 Bearing Stress on Bedrock


For Conventional Abutment
CL Base (Footing) LS
DLv, LLv

XDL, x LL

W1

PLS

V1

Xv1

Ph

Xw1

H/3

W2

e
R
B/2
B 2e
B

Step 1: Calculate the eccentricity about point C, ec,, where:


Mo = sum of moments of factored overturning forces, acting about point C:

M o Ph

H
H
W1 X w1 PLS ,h DLV X DL LLv X LL
3
2

Mr = sum of moments of factored resisting forces about Point C:

M r V1 X V 1 LS L X LS
V = sum of factored vertical forces acting on the footing and wall:

March 2014

5-32

CHAPTER 5 - SUBSTRUCTURES

V V1 W1 W2 DLv LLv LS L
and,

ec

ec

Mo Mr
V
PH

H
H
V1 X V ! LS L X LS W1 X W 1 PLS ,h
3
2
V1 W1 W2 DLv LLv LS L

Step 2: The factored vertical stress should be calculated assuming a linearly


distributed pressure over an effective base area shown in the figure above. If the
resultant is within the middle 1/3 of the base, the maximum and minimum factored
vertical stress is calculated as follows:

v max

v min

e
V
1 6 c
B
B

e
V
1 6 c
B
B

If the resultant is outside of the middle 1/3, of the base, i.e., if B/6, vmin will drop to
zero, and as e increases, the portion of the heel of the footing which has zero vertical
stress increases.

v max

v min

2 V

B
3 ec
2

Step 3: Comparevmax to the factored bearing resistance, q r, provided in the


Geotechnical Report. The maximum factored bearing stress should be less that the
factored bearing resistance.

v max

bc Rn R f

where:
vmax = maximum factored vertical stress (ksf)
bc = bearing resistance factor (dim)
Rn = nominal bearing resistance (ksf)
Rf = factored bearing resistance (ksf)

March 2014

5-33

CHAPTER 5 - SUBSTRUCTURES
Note: The case shown for this procedure is the construction load with full backfill and
live load surcharge on the approach. For other load combinations the appropriate
loads must be included in the analysis.

F.

Sliding

Failure by sliding should be investigated for all abutments founded on


spread footings bearing on soil or bedrock. Passive earth pressure
exerted by fill in front of the footing is neglected in consideration that soil
may be removed during future construction. Refer to Section 3.6.9
Passive Earth Pressure Loads for guidance. The factored resistance
against failure by sliding of abutments and walls on spread footings shall
be calculated as described in Section 5.3.8 and LRFD 10.6.3.4.
Resistance factors for sliding of spread footings at the strength limit state
are provided in Table 5-3.
The coefficient of friction for sliding should be as shown in Table 3-3 for
the appropriate soil type under the footing. For footings on bedrock, the
Geotechnical Designer will provide a coefficient of friction for sliding,
based upon the bedrock characteristics.

March 2014

5-34

CHAPTER 5 - SUBSTRUCTURES

Procedure 5-3 Eccentricity and Sliding Check for Conventional


Abutment on Spread Footing
LS

W1

V1
PLS

XV1
Xw1

Ph

W toe

L
Xw2

W2

eo

QT

R = V = resultant of factored forces at the base of


footing

eo = eccentricity of resultant, calculated


around point O (toe of footing)

Step 1: Calculate the eccentricity about Point O in the Figure above to locate the
resultant force R. Forces and moments resisting overturning are to be considered
negative, and the maximum load factors should be used (Table 5-4)
Mo = sum of moments of factored overturning forces acting about Point O:

M o Ph

H
H
P LS
3
2

Mr = sum of moments of factored resisting forces acting about Point O:

M r V1 xV 1 W 1x w1 W2 sw2
V = sum of factored vertical forces acting on footing and wall, as defined in the
Figure above.

March 2014

5-35

CHAPTER 5 - SUBSTRUCTURES

V V1 W1 W2
Step 1: Check eccentricity (overturning) about Point O:

eo

Mo Mr
V

For footings on soil, the location of the resultant force shall be within the middle twothirds (2/3) of the base width. For footings on bedrock, the location of the resultant
force shall be within the middle nine-tenths (9/10) of the base width. For footings
subjected to biaxial loading, these eccentricity requirements apply in both directions.
Step 2: Compare the factored resistance to sliding to the factored applied horizontal
loads. The factored resistance to sliding should be greater than the factored applied
horizontal loads:

Rn V tan
R f Rn s Q
where:
Rn = Nominal sliding resistance between soil and foundation (kips)
friction angle between the footing base and the soil (refer to Table 3-3 or LRFD
Table 3.11.5.3-1.)
s = resistance factor for shear resistance between the soil (or rock) and foundation
Q = factored horizontal applied loads
Note: The load combination shown for this strength limit state is Strength I-a, which
does not consider superstructure dead loads (DC and DW) and vehicular live loads
(LL) For other load combinations the appropriate loads and load factors must be
included in the analysis.

G. Eccentricity
Abutments and walls on spread footings should be designed to resist
overturning which results from lateral and eccentric vertical loads. The
eccentricity should be evaluated as shown in Procedure 5-3. The
location of the resultant of the reaction forces of at the strength limit
state, based on factored loads, shall be within the middle two-thirds (2/3)
of the footing width for footings on soil or the middle nine-tenths (9/10) of
the footing width for footings on rock.
If construction loading is critical, the backfill height may be restricted until
the superstructure or other parts are constructed.

March 2014

5-36

CHAPTER 5 - SUBSTRUCTURES

5.4.1.12 Abutments Supported on Pile Foundations


Piles should be designed in accordance with the requirements of Section
5.7 Piles.
For pile supported abutments, the factored load combination causing the
maximum and minimum compression in the piles should be determined,
and the resulting pile reactions and pile stresses determined. The
maximum factored axial pile load should not exceed the lesser of the
factored geotechnical resistance and factored structural resistance for a
single pile. In accordance with LRFD Article 6.5.4.2, the factored pile loads
should not exceed the factored structural resistance using the resistance
factors provided in 5.7.2 H-Piles and 5.7.5 Steel Pipe Piles. If greater
loads result, more piles, or larger piles, should be considered.
For the Service Limit State, the unfactored lateral pile loads for H-piles
should not exceed the lateral loads resistances specified in 5.7.2.2
Load combinations that do exceed the lateral load limits established for
the service limit state should be evaluated by the Geotechnical Designer
by means of a project-specific pile lateral load analysis using LPILE
software. The maximum lateral loads for all piles other than steel H-piles
should be evaluated by the Geotechnical Designer. Buckling analyses of
piles should be performed by the Structural Designer. Piles should also
be checked for resistance against combined axial loads and flexure per
LRFD 6.15 and 6.9.2.2. Pile resistance should be determined for
compliance with the LRFD interaction equation.
Where abutments are required in water channels, the bottom of seal
should be a minimum of 2 feet below the calculated scour depth from the
check flood for scour. Where the calculated scour depth is significant, the
Designer may consider designing the deep foundation elements for an
unsupported length. The unsupported length should be the vertical
distance from the bottom of the seal to the check flood scour depth. In
designing deep foundation elements for an abutment with an unsupported
length, a complete analysis of the foundation should be performed using
actual loading and soil conditions.
Vertical and horizontal movement criteria for abutments supported by pile
foundations should be developed consistent with the function and type of
structure. The effect of lateral squeeze in the pile-supported abutments
should be considered by the Geotechnical Designer, if applicable. Refer
to Sandford, October 1994.

March 2014

5-37

CHAPTER 5 - SUBSTRUCTURES

5.4.1.13 Bridge Seat Dimensions


As a minimum, the bridge seat dimensions should meet the requirements
of LRFD Article 4.7.4.4. Otherwise, for bridge seats supporting steel
superstructures exceeding 200 feet, use a minimum of 2 feet between the
centerline of bearings and the face of breastwall and a minimum of 2-3
between the centerline of bearings and the face of backwall. The masonry
plate of the bearings should be no closer to the face of breastwall than 3
inches and should clear the face of backwall by at least 2 inches. For
steel superstructures between 100 and 200 feet use a minimum 3 foot
bridge seat. For steel superstructures less than 100 feet, the bridge seat
dimensions should be large enough to accommodate the bearing masonry
plate and the previous clearance dimensions. For major steel structures,
all precast concrete structures, and structures with skews exceeding 45,
the bridge seat dimension should be determined based upon the project
requirements.
All bridge seats, regardless if protected from roadway drainage by sealed
bridge joints, should be concrete pedestal type with a minimum width
along the centerline of bearing of 3 feet. The clear distance between the
ends of bearing masonry plates and the ends of concrete pedestals
should be at least 6 inches. The bridge seat between concrete pedestals
should be sloped downward toward the face of breastwall at a slope of at
least 15%.
Top of abutment backwalls should be 1-6 wide, excluding the 6 inch
approach slab seat, except when the concrete superstructure slab extends
over the top of the backwall and the back of the backwall is battered. In
that case, the backwall should be 1-6 plus the effect of the batter.
5.4.2 Integral Abutments
5.4.2.1

Introduction

There are two categories of integral abutments: (1) full integral abutments,
where the bridge beams are rigidly cast into an end diaphragm and (2)
integral with hinge abutments, where butted boxes or voided slabs are
connected to the abutment with dowels.
Integral abutment bridges (IABs) should be evaluated for use on all bridge
replacement projects. MaineDOT most commonly uses 4 piles for each
integral abutment substructure unit and traditionally uses the following
piles:

March 2014

HP 10x42

HP 12x53
5-38

CHAPTER 5 - SUBSTRUCTURES

HP 14x73

HP 14x89

Design is not limited to these piles. If the Structural Designer elects to use
a pile not listed, the appropriate design analysis must be conducted.
Although HP 14 x 73 pile flanges are non-compact and do not meet the
slenderness requirements of LRFD 6.9.4.2, Designers can account for pile
slenderness in the design process, and this pile size should still be
considered for pile supported integral abutments.
5.4.2.2

Loads

Analysis and design of integral abutment substructures will be in


accordance with AASHTO LRFD, and include structural design and
analysis of reinforced concrete abutments and wings, global stability of the
channel slope with abutment, and pile design. Load combinations are
presented in Section 5.4.1.2. Additional appropriate load combinations
that investigate the effects of thermal gradients and abutment
displacement may be required in accordance with LRFD Section 3.
5.4.2.3

Historical ASD Design Practice and Bridge Lengths

Commentary: Design of integral abutment bridges has evolved over the years
as transportation departments have gained confidence with the system. Bridge
lengths have gradually increased without a rational design approach.
Tennessee, South Dakota, Missouri and several other states allow lengths in
excess of 300 feet for steel structures and 600 feet for concrete structures.
Thermally-induced pile head translations in bridges with the lengths stated
above will cause pile stresses which exceed the yield point. Research
performed during the 1980s (Greimann, et. al.) resulted in a rational design
method for integral abutment piles, which considers the inelastic redistribution
of these thermally induced moments. This method is based upon the ability of
steel piles to develop plastic hinges and undergo inelastic rotation without local
buckling failure. This method is not recommended for concrete or timber piles,
which have insufficient ductility.
Past practice was based on evaluation of the four steel piles most commonly
used by MaineDOT and maximum bridge lengths and maximum design pile
load design guides were developed based upon the Greimann research. The
pile were evaluated as beam-columns without transverse loads between their
ends, fixed at some depth and either pinned or fixed at their heads.

Greimann, et. al., developed design criteria by which the rotational


demand placed upon the pile must not exceed the piles inelastic rotational
capacity. The following system variables affect the demand:
March 2014

5-39

CHAPTER 5 - SUBSTRUCTURES

Soil type

Depth of overlying gravel layer

Pile size

Pile head fixity

Skew

Live load girder rotation

In order to simplify the design, past practice assumed that piles would be
driven through a minimum of 10 feet of dense gravel. Material below this
level has very little influence on pile column action. It was also assumed
that the live load girder end rotation stresses induced in the pile head do
not exceed 0.55 Fy (which provides a known live load rotational demand).
Based upon the above assumptions and the piles inelastic rotational
capacity, the maximum pile head translation, (in inches) was established
for each of the four piles. Based on allowable stress design, the maximum
bridge lengths historically were as follows:

MaxBridgeL ength ft

4 in
0.0125 for steel bridges

MaxBridgeL ength ft

4 in
0.075 for concrete bridges

Maximum bridge lengths vary from 70 feet to 500 feet for some piles. The
past practice for maximum bridge lengths was 200 feet for steel and 330
feet for concrete. FHWA allows maximum bridge lengths of 300 feet for
steel bridges, 500 feet for cast-in-place concrete bridges, and 600 feet for
prestressed or post tensioned concrete bridges (FHWA Technical
Advisory, January 28, 1990). Refer to BDG 5.4.2.6 for current bridge
length limits.
5.4.2.4

Pile Design

A. Pile Loads
Piles should be modeled and evaluated as either fixed at the pile head
for fully integral abutments (bridge beams are rigidly cast into an end
diaphragm) or as pinned for integral abutments with hinge, such as the
case when butted boxes or voided slabs have dowel connections to the
abutment.

March 2014

5-40

CHAPTER 5 - SUBSTRUCTURES

Piles for full integral and integral with hinge abutments shall be designed
to resist all vertical superstructure dead and live loads, abutment and
pile dead loads, live load girder rotation moments, lateral displacements,
live load impact and moments caused by superimposed dead loads and
live loads, as appropriate for the type of integral abutment.
Until the behavior of integral abutments with hinged connections to the
superstructure is better understood, the pile design criteria for that type
of integral abutment may assume that the moment at the top of the pile
is zero, and that there is no moment from either the superstructure or
earth loads.
The effect of thermal displacements and moments on piles can be
investigated by running LPILE software.
Secondary thermal forces only need be considered for multi-span
structures only.
Appropriate load combinations and load factors should be determined
per LRFD 3.4.1.
For the strength limit state analysis, design of the piles should consider
the factored structural pile resistance, Pr, the factored structural flexural
resistance, pile unbraced length, pile moments, the interaction of
combined axial and flexural load effects, the structural shear resistance
and the factored geotechnical resistance.
For service limit state evaluations, if piles will be driven to practical
refusal in bedrock, settlement will not be a concern. However, all
designs should consider horizontal movement, overall stability and scour
for the design flood event.
B. Resistance Factors for Integral H-Piles
Pile will typically be end bearing on bedrock. For the strength limit state,
use the following resistance factors:
o Use c = 0.50 for axial resistance in compression and subject to
severe pile driving condition; this condition should be assumed
when analyzing the lower portions of the pile
o Use c = 0.60 for axial resistance in compression under good
driving conditions; this condition should be assumed when
analyzing the upper portion of the pile
o For combined axial and flexural resistance in the upper zone of
pile, use:

March 2014

5-41

CHAPTER 5 - SUBSTRUCTURES

c = 0.70 for axial resistance

f = 1.00 for flexural resistance

C. Design Steps
The following steps should be followed during design of piles supporting
full integral abutments, for the strength limit state:
1. Determine the foundation displacements, and the load effects (Pu
and Mu) from the superstructure and substructure designs.
2. If applicable, determine the magnitude of scour.
3. Select preliminary pile size:
a. Determine the factored applied superstructure vertical dead
and live load (Pu) distributed to each pile
b. Select the steel pile strength
c. Select pile orientation; typically weak axis bending
d. Determine resistance factors (c and f) for the structural
strength in the upper and lower zones of the pile.
e. Determine the maximum, required nominal axial pile
resistance, Pu/f
f. Estimate an initial pile area using the approximation
Ru
A.s
0.80 Fy
This approximation is based on weak axis bending and an
assumed unbraced length of 15 feet based on typical integral
abutment pile deflection and moment with depth curves.
Select a pile size with an area As or greater.
4. Determine the pile unbraced length and maximum moment at the
top of the pile by running LPILE software for the design
displacement from Step 1, Pu, and live load rotation
5. Determine if the applied moment on the pile will cause pile head
plastic deformation by using the Interaction of combined axial and
flexural load effects on a single pile (LRFD 6.9.2.2)
a. Obtain the unbraced lengths of the top and lower segments of
the pile and calculate the column slenderness factor () for
each segment. (LRFD 6.9.4.1)
b. Determine K values for the top and bottom of the pile per
LRFD Table C4.6.2.5-1

March 2014

5-42

CHAPTER 5 - SUBSTRUCTURES

g. Calculate the nominal and factored structural pile resistance


Pn, per LRFD 6.9.4.1 using the values
h. Compare the ratio of Pu to the structural resistance in the
upper portion of the pile the pile size should be such that the
ratio is not less than 0.20.
i. Determine the nominal and factored flexural resistance about
H-Pile weak axis, (LRFD 6.12.2.2)
j. Calculate the moment that will cause a plastic hinge at the top
of the pile (Mp)
k. If the applied moment exceeds the moment that would cause
a plastic hinge, a plastic hinge forms, and the moment that can
be applied cannot exceed that moment (Mp)
6. For fixed head piles, run a second LPILE analysis with
displacement and plastic moment (Mp) as load conditions and Pu,
and calculate new unbraced lengths from the moment with depth
curve.
a. Repeat steps 5.a. through 5.d., above
b. If the pile size is such that the ratio of Pu to structural
resistance exceeds 0.2, check the upper zone of the pile with
the interaction equation of LRFD 6.9.2.2. If a plastic hinge
forms at the top of the pile, the K value of the upper segment
(that portion between the top of the pile and the first inflection
point on the moment vs. depth curve) changes from 1.2, for a
pinned condition, to 2.1, for a free condition at the top. With
the new K value repeat Step 5, and check the interaction
equation for pile overstress.
7. Because the piles have weak axis orientation and the flanges
resist the shear as opposed to the web, check the maximum
shear from the LPILE output to the structural shear resistance
per AISC G7.
8. Check that the maximum factored applied pile load does not
exceed the factored geotechnical pile resistance or pile drivability
resistance (LRFD 10.5.5.2.3 and 10.7.3.13) provided in the
Geotechnical Design Report.
5.4.2.5

Pile Length Requirement

A. General Requirements
Piles may be end bearing or friction piles. In order to obtain the pile
behavior associated with the equivalent length, piles should be installed
1 to 5 feet beyond the pile length required to achieve fixity. The practical

March 2014

5-43

CHAPTER 5 - SUBSTRUCTURES

depth to pile fixity is defined as the depth along the pile to the point of
zero lateral deflection.
A minimum pile length of 10 feet is recommended, however soil
conditions and loading conditions may require additional pile embedment
to achieve fixity. Additional embedment length may be required for the
use of friction piles. Also, axial loads may govern and additional
embedment length may be required in order to achieve the factored
design axial load with appropriate resistance factor applied. For pile
lengths less than 14 feet, consideration should be given to the pile
translating as a column and the pile tip walking. More vigorous driving
shoes designed to properly seat piles and hold the pile and point in place
are available. Refer to paragraph B. Short Pile Usage Guidelines,
below.
If site-specific soil properties and loading conditions exist, an evaluation
of minimum embedment length to achieve fixity using LPILE software or
the Davisson and Robinson equation in LRFD 10.7.4.2 is recommended.
Consult the Geotechnical Designer for these analyses.
Piles should be driven with their weak axis perpendicular to the
centerline of the beams, regardless of skew. Refer to Section 5.7.2 HPiles for additional design requirements.
When scour is anticipated, the minimum pile length should be provided
beyond the depth of computed scour.
B. Short Pile Usage Guidelines
The MaineDOT and the University of Maine at Orono (UMaine) have
investigated the performance of integral abutment bridges at sites with
shallow bedrock and have instrumented and monitored Nash Stream
Bridge in Coplin Plantation, Maine, (Hartt, et. al., 2006 and Delano, et.
al., 2005)). Preliminary evaluation of the field data from the research
study indicate that integral abutment bridges with short steel piles (14
feet or less) may not develop fixity but perform adequately and do not
experience stresses larger than those seen by longer piles. The shortest
pile instrumented by the researchers was a 14-foot long H-pile.
To accommodate integral abutment piles at sites with shallow bedrock,
the following design features are recommended:
o In consideration of (a) the consequences of scour and pile
exposure, (b) the need to limit pile tip movement, and (c)
obtaining pile behavior associated plastic stress redistribution and
inelastic rotation in the pile, a minimum pile length of 10 feet is
recommended. This recommendation is based on finite element
analyses and limited field data from the UMaine studies (Delano,
March 2014

5-44

CHAPTER 5 - SUBSTRUCTURES

et. al. 2005 and Hartt, et. al. 2006). If the depth to bedrock is so
shallow that 10 feet of embedment in soil cannot be achieved,
piles should be installed in bedrock sockets to provide the
minimum 10-foot pile length recommended. If a fixed condition at
the pile tip is desired, the bottom 6-inches of the rock sockets
should be tremie-filled with concrete. However, the UMaine
research indicates some rotation at the pile tip is acceptable.
o Short piles supporting integral abutments should be designed in
accordance AASHTO LRFD criteria and checked for pile tip
movement by conducting a LPILE analysis, or as described in
the design example found in Appendix B of Technical Report ME
01-7 (Delano, et. al. 2005), and Chapter 5 of that report.
Achievement of an assumed pinned condition at the pile tip
should also be confirmed with an LPILE analysis.
o Since the abutment piles will be subjected to lateral loading, the
piles should be analyzed for combined axial compression and
flexure resistance as prescribe in LRFD Articles 6.9.2.2 and
6.15.2 and checked for compliance with the interaction equation.
An LPILE analysis is recommended to evaluate the soil-pile
interaction with factored axial loads, moments and pile head
displacements applied.
o Driven piles should be fitted with special driving points to improve
penetration into bedrock and improve friction at the pile tip to
support a pinned pile tip assumption.
o The stream velocity should be low and there should be low
potential for removal of any dams, scour action, wave action,
storm surge and ice damage. This is to ensure the long-term
integrity of the bridge approach fills and riprap abutment slopes,
which provide the only lateral support to pile groups.
o Minimum 1.75H:1V slopes in front of integral abutment pile
groups should be protected with riprap over an erosion control
geotextile or concrete slope protection.
5.4.2.6

Maximum Bridge Lengths

The criteria for the maximum bridge lengths provided in Table 5-5 are
based on the following assumptions:

March 2014

Steel H-piles are used with their webs oriented normal to the
centerline of the bridge (longitudinal translation about the weak
axis).

5-45

CHAPTER 5 - SUBSTRUCTURES

The piles are driven through gravels or through clays with a


minimum of 10 feet of gravel overburden.

For skews greater than 20, abutment heights are <12 feet and
pile spacing is < 10 feet.

Total thermal movement is 1-1/4/100 feet bridge length for steel


structures and 3/4/100 feet bridge length for concrete
structures (FHWA Technical Advisory, January 28, 1990).

Factored pile loads do not exceed the factored compressive


structural pile resistance, the factored flexural pile strength and
the factored geotechnical and drivability resistance of the pile
section.

Steel H-Piles are made of Grade 50 steel.

Bridge lengths in excess of the limitations below may be used with the
approval of the Engineer of Design when special design features are
provided.
Table 5-5 Recommended Maximum Lengths for Fully Integral
Abutment Bridges (feet)
Pile Size
Piles per 5.4.2.1 with
fully fixed heads
5.4.2.7

Steel
300

Skew 20
Concrete
500

Best Practices for Moderate to Long Span IABs

The following best practices should be considered as design features for


moderate to long span integral bridges, defined as integral steel bridges
longer than 200 feet and concrete bridges longer than 330 feet:

March 2014

Only straight stringers/beams should be used on long span


IABs.

The annual thermal cyclic movement of the IAB abutments


results in the development of a settlement trough adjacent to
each abutment as backfill soil slumps downward and toward the
abutment in its winter position. To prevent the settlement of the
pavement structure, approach slabs must be included in the
design of moderate to long span IAB structures, to span over
the void created by the settled soil.

5-46

CHAPTER 5 - SUBSTRUCTURES

March 2014

Provide 2 layers of polyethylene sheets, or other bond breaker,


under the approach slab to minimize friction against horizontal
movement. Many States recommend two layers of 4 to 6 mil
thick polyethylene sheets.

Consider pavement expansion joints to reduce distress of the


approach pavement, caused by the thermal cyclic movement of
the abutments and the approach slabs. Recommended cycle
control joints systems that employ a combination of asphaltic
plugs, asphalt impregnated fiber board, and sleeper slabs at the
end of the at-grade approach slabs, or at the end of the
abutment (in the case where the slab is buried).

A bridge with a total length in excess of 300 feet will have larger
movement demands. If the anticipated abutment movements
are in excess of 1.0 inch, consider strong axis pile orientation to
prevent a plastic hinge under weak axis bending.

Approach slabs should also be positively attached to the


abutment to prevent slabs from walking off corbels during
annual thermal movements of the abutment.

Pavement geotextiles can be used to add tensile strength to


pavement over the abutment backwall.

Provide adequate drainage of the abutment backfill to prevent


damage due to frost action and piping of the backfill material.

Bridge abutments with movements in excess of 1 inch may


require a higher level of pile analysis to consider all applicable
forces and moment demands, including thermal, skew effects
and deflections of the superstructure. A dedicated check of pile
capacity for combined axial loading due to dead and live load
and bending stresses due to thermal superstructure movement,
using LPILE software may be required.

Pre-auger to a depth of 10 feet for the top portion of piles and


then fill the hole with a non-compacting backfill material, such as
underdrain backfill Type C. This creates a hinge effect in the
substructure and has the effect of reducing the lateral soil
stiffness by increasing the depth to fixity and reducing bending
moment stress in the pile.

Long-span integral bridges receive significant support from the


embankments, and therefore, they only should be built in
conjunction with stable approach embankment foundation soils.

To mitigate excessive earth pressures, limit abutment heights.


5-47

CHAPTER 5 - SUBSTRUCTURES

Avoid abutments of differing height; as such a practice may


promote unequal movements at the two abutments.

Select a span arrangement and bearing types that result in


approximately equal movements at each abutment.

As a result of the soil movement, the summer lateral earth


pressures tend to increase over time as the soil immediately
adjacent to each abutment becomes increasingly wedged in.
This phenomenon of soil wedging and long-term buildup of
lateral earth pressures is referred to as ratcheting. To avoid
potential problems, abutments should be designed for full
passive pressure using Coulomb Theory.

Limit the use of long span integral abutments to bridges with


skews less than 20 degrees to minimize the magnitude and
lateral eccentricity of potential longitudinal forces.

Make wingwalls as small as practical to minimize the amount of


structure and earth that have to move with the abutment.

Configure wingwalls to minimize resistance to abutment


movement.

5.4.2.8

Abutment Details

Typical integral abutment details for steel and concrete superstructures


are shown in Figure 5-2 and Figure 5-3, respectively. For steel
superstructures, fixed head integral abutments are preferred but pinned
head abutments are allowed.

March 2014

5-48

CHAPTER 5 - SUBSTRUCTURES

Figure 5-2 Fixed Pile Head, Full Integral Abutment Details-Steel


Superstructures

March 2014

5-49

CHAPTER 5 - SUBSTRUCTURES

Figure 5-3 Integral Abutment with Hinge and Full Integral Abutment Details
Precast Superstructures

March 2014

5-50

CHAPTER 5 - SUBSTRUCTURES

5.4.2.9

Alignment

Curved bridges are allowed, provided the stringers are straight. Beams
should be parallel to each other. All substructure units should be parallel
to each other.
The maximum vertical grade between abutments is limited to 5%.
5.4.2.10 Superstructure Design
No special considerations should be made for integral abutment designs.
Fixity at the abutments should not be considered during beam/girder
design.
When selecting span ratios for multi-span bridges, consideration should
be given to providing nearly equal movement at each abutment.
5.4.2.11 Abutment and Wingwall Design
Design abutment and wingwall reinforcement for the passive earth
pressure (Pp) which results on the back face of the wall when the bridge
expands. Refer to Section 3.6.6 Coulomb Passive Lateral Earth Pressure
Coefficient (Kp) and Table 5-4 for the passive earth pressure load factor
(EH).
Design bars for the backwall for full passive pressure due to the abutment
backfill material. The backwall acts as a continuous horizontal beam
supported on the piles, i.e., with spans equal to the girder spacing.
Design the bars for 1) the maximum factored shear due to the factored
passive earth pressure and, 2) flexure due to the moment from the
factored passive soil pressure Determine the passive pressure Pp acting
on the full height of the abutment backwall (Habut) from the bottom of the
approach slab to the bottom of the abutment/pile cap. The passive
pressure acts in a triangular pressure distribution:

1
2
Pp soil Habut k p
2
Design for a factored moment equal to:

Mup EH

Pp l

A load factor for passive earth pressure is not specified in LRFD. Use the
maximum load factor for active earth pressure, EH = 1.50.
March 2014

5-51

CHAPTER 5 - SUBSTRUCTURES

Design the abutment wall top and bottom horizontal bars for vertical loads,
considering the wall to be a continuous beam with piles as supports.
Wingwalls should preferably be straight, cantilevered extension wings not
to exceed 10 feet in length. Design wingwall reinforcement for the passive
earth pressure (Pp) which results on the back face of the wall when the
bridge expands, using the Coulomb passive earth pressure state and a
passive earth pressure load factor (EH) of 1.5. The use of flared
wingwalls may be considered at stream crossings where the alignment of
the stream would make in-line walls subject to scour. Piles should never
be placed under wingwalls that are integral with the abutment stem.
Generally the design is controlled by the horizontal bending in the
wingwall at the fascia stringer caused by large passive pressures bending
the wingwall.
Because of the high bending moments due to passive pressure in
wingwalls 10 feet or longer, it may be necessary to support longer
wingwalls on their own foundations, independent of the abutments. A
flexible joint must be provided between the wingwalls and the backwall.
U-wingwalls cantilevered from the abutment stem should only be
considered to address right-of-way or wetlands encroachment. Uwingwalls should be no longer that 10 feet and tapered to reduce earth
pressures. If an approach slab must extend to a U-wingwall, use a 2 inch
joint with filler to separate the slab and the wall.
Developing full passive earth pressure requires that wall rotation, i.e. the
ratio of lateral abutment movement to abutment height (y/H), exceeds
0.005. If the calculated rotation is significantly less than that required to
develop full passive pressure, the Designer may consider using the
Rankine passive earth pressure case, which assumes no wall friction. For
the passive earth pressure case, wall friction acts downward against the
passive wedge and increases passive pressure in the Coulomb state.
5.4.2.12 Approach Slabs
In addition to the requirements of Section 5.4.4, approach slabs should be
used when integral bridge lengths exceed 80 feet for steel structures and
140 feet for concrete structures.
Provisions for movement between the approach slab and approach
pavement is not necessary until bridge lengths exceed 140 feet for steel
structures and 230 feet for concrete structures. Approach slabs below
grade should be attached to the abutment. For at grade approach slabs,
consideration should be given to the installation of an expansion device
between the approach slab and the abutment. Refer to recommendations
for approach slabs for moderate to long span integral bridges in Section
5.4.2.7.
March 2014

5-52

CHAPTER 5 - SUBSTRUCTURES

5.4.2.13 Drainage
The area behind integral abutments should be backfilled with granular
borrow for underwater backfill. A proper drainage system as described in
Section 5.4.1.9 should be provided to eliminate hydrostatic pressure and
control erosion of the underside of the abutment embankment slope
protection. A drainage system is of great importance when there is
potential for a perched or high groundwater condition, when the bridge is
located in a sag curve, when the bridge is located in a cut section with
saturated subgrade, or when there is significant pavement water runoff to
side slopes. In these situations, consideration should also be given to
backfilling integral abutments with gravel borrow or aggregate subbase
course - gravel.
5.4.2.14 Scour
The Designer should ensure the stability of the structure for anticipated
scour, as defined by LFRD 2.3.11. This may require driving the piles
deeper than what is required by geotechnical criteria. The minimum pile
length should be provided beyond the depth of computed scour for the
check flood for scour.
5.4.2.15 Integral Abutment on Spread Footing Design
Spread footing abutments may be used only if designed and detailed as a
semi-integral bridge abutment. Refer to Section 5.4.3 Semi-Integral
Abutments.
5.4.3 Semi-Integral Abutments
A semi-integral bridge is defined as a single span or multiple span continuous
deck-type bridge with rigid non-integral abutment foundations, and with a
movement system composed primarily of reinforced concrete end-diaphragms,
backfill, approach slabs, and movable bearings located in horizontal joints at
the superstructure/abutment interface (TRB, 1996).
A semi-integral abutment bridge is characterized by:

Elimination of expansion joints in the deck and roadway

The superstructure backwall (end diaphragm) is not connected to the


abutment, but moves along a bearing and horizontal joint below
ground

Thermal movement is accommodated by expansion bearings and a


small vertical gap between the end diaphragm and the abutment

March 2014

5-53

CHAPTER 5 - SUBSTRUCTURES

The abutments are typically supported on spread footings or multiple


rows of piles

Semi-integral abutments should typically be designed for active earth pressure


over the rigid abutment height and a uniform pressure distribution due to the
height of soil behind the superstructure. The superstructure backwall should
typically be designed for full passive pressure only. In designing for active
pressure, a Rankine active earth pressure coefficient, Ka, is recommended.
Semi-integral bridge design is still considered experimental, and must receive
approval from the Engineer of Design during the preliminary design phase as
a design exception.
Research findings have resulted in TRB design recommendations that include
the following:

Utilization of attached approach slabs and return wingwalls to lock


the superstructure into the backfill

Deliberate construction of an air space below the end diaphragms to


prohibit an undesirable shift in the end reaction location

5.4.4 Approach Slabs


Approach slabs should be used on collectors and arterials, where:

the design hour volume (DHV) is greater than 200,

abutment heights (bottom of footing to finish grade) are greater than


20 feet, or,

poor soil conditions are encountered and settlement is anticipated in


the vicinity of the abutment.

Additional requirements for the use of approach slabs on integral abutment


bridges are provided in Section 5.4.12.
Approach slab seats should be 6 inches wide and specified to have a
roughened surface. Approach slab seat dowels should not be used except on
integral abutments as discussed in Section 5.2.4.12. Approach slab seats
should be a minimum vertical distance of 2-9 from the roadway surface. If
the backwall is very high, the Structural Designer may elect to make an
optional horizontal construction joint at the approach slab seat elevation.
When a structural approach slab is specified, reduction, but not elimination, of
the vehicle surcharge loads may be considered per LRFD 3.11.6.5.

March 2014

5-54

CHAPTER 5 - SUBSTRUCTURES

5.5

Piers

5.5.1 Mass Piers


Mass piers are intermediate vertical supports, which extend from the
foundation, either a spread footing or deep foundation, to a pier cap, which
supports the superstructure. The primary functions of pier are:

Support dead loads, live loads and other loads from the
superstructure

Support its own weight and other loads acting directly on the pier

Transmit all loads to the underlying foundation

The connection between the pier and the superstructure may be pinned, fixed,
or free. Mass piers are typically constructed from reinforced concrete, but may
be precast. Mass piers may consist of gravity, solid wall, single-column, or
multiple-column piers. Single-column and multiple-column piers are usually
designed in a hammerhead configuration at the pier cap.
5.5.1.1

Pier Selection Criteria

Selection of the mass pier configuration is based on the following factors:

Loading conditions

Skew

Slenderness, with respect to buckling

Aesthetics

Likelihood of debris. The use of multiple-column piers in areas


where floating debris may lodge between columns should be
avoided.

5.5.1.2

Load Combinations and Load Factors

Mass piers should be designed in accordance with AASHTO LRFD,


including, structural design of reinforced concrete and geotechnical
analysis and design, such as bearing capacity, sliding, and eccentricity
(overturning). Piers should be designed and proportioned to resist all
applicable load combinations specified in LRFD Articles 3.4.1 and 11.5
and as outlined in Chapter 3.

March 2014

5-55

CHAPTER 5 - SUBSTRUCTURES

The following load combinations should be considered as a minimum for


geotechnical analysis:

Strength I-construction. Strength Limit State I with the


exception that bridge superstructure, or a portion of that, and
vehicle live loads are neglected. Any anticipated staged
construction loading should be investigated. Load factors for
the dead load or other components shall not be less than 1.25.
Live load surcharge is included to account for construction
equipment live loading during structure erection, and a
construction load factor of 1.5 should be assumed.

Strength I: Normal vehicular use without wind: dead load, all


applicable live load combinations, impact; braking force (for one
and two lanes) centrifugal forces, static water pressure,
buoyancy and stream pressure. For Strength 1, the minimum
and maximum permanent load factors are selected to create the
greatest force and moment effects for the mode of stability
being investigated. .

Strength III: Load combination relating to high wind velocity


(100 mph) without vehicular live load: dead load, earth pressure,
if applicable; buoyancy; stream flow pressure; wind; wind on live
load; and longitudinal force from thermal displacements.
Minimum and maximum load factors for permanent loads (p)
are selected to produce the extreme force or moment effect for
sliding, eccentricity or axial loading analyses.

Strength IV: Load combination relating to very high dead load


to live load force effect ratios exceeding about 7.0. Minimum
and maximum load factors for permanent loads (p) are be
selected to produce the extreme force or moment effect for
sliding, eccentricity or axial loading analyses.

Strength V: Load combination relating to the bridge exposed to


55 mph wind velocity with live loads: dead load; live load plus
impact; centrifugal force; earth pressure; buoyancy and stream
flow pressure. . Minimum and maximum load factors for
permanent loads (p) should be selected to produce the extreme
force or moment effect for sliding, eccentricity or axial loading
analyses.

Service I: Normal vehicular use of the bridge with a 55 mph


wind load. All loads are taken at their unfactored values.

Debris loading shall be accounted for in water pressure loads by a 25%


increase in the exposed surface area of the pier.

March 2014

5-56

CHAPTER 5 - SUBSTRUCTURES

A Maine-modified Strength Limit State analysis should be performed that


includes in the ice pressures of past practice, specified in Section 3.9 Ice
Loads, with the appropriate resistance factors applied to the pier
components. The Strength Limit State that produces the extreme force or
moment should be selected.
Where piers are to be designed to resist earthquake forces, collisions by
roadway or rail vehicles, vessel collision or ice, the structures should be
evaluated for the following additional limit states:

Extreme Event I Load combination including earthquake


forces, using permanent load factors, p, which produce the
greatest load and moment effects for the mode of stability being
analyzed.

Extreme Event II Dead load; live load; buoyancy; static water


pressure; stream flow pressure; ice pressure; vessel impact and
vehicle or railway impact, using permanent load factors, p,
which produce the greatest load and moment effects for the
mode of stability being analyzed.

For Extreme Event II apply ice force effects, and vessel, vehicle and
railway collision forces one at a time since the joint probability of these
events is extremely low.
The ice pressures for Extreme Event II shall be applied at Q1.1 and Q50
elevations as defined in Section 3.9 Ice Loads with the design ice
thickness increased by 1 foot and a load factor of 1.0.
The critical load conditions for the evaluation of foundation bearing
capacity, overturning (for pile foundations assess uplift loading of piles),
eccentricity, and sliding (lateral loading for deep foundations) are those
combinations of minimum or maximum loads and moments which
produces the maximum force or moment effect.
With regards to vehicular live load (LL and IM) lane placement is important
and multiple presence factors (MPF) are applicable. Impact forces should
only be applied to truck or tandem loads:

IM = 0.33 for cap and stem

IM = 0 for buried footings

In consideration of the potential deflections due to bending of a pier about


its weak (transverse) axis may result in magnification of the longitudinal
moments on the pier, the Designer should compute longitudinal moment
magnification factors for each load combination and Strength Limit State

March 2014

5-57

CHAPTER 5 - SUBSTRUCTURES

based on the factored loads and pier stiffness. The Moment


Magnifications Factors are provided in LRFD 4.5.3.2.2.
5.5.1.3

General

The designer should estimate the load combinations which could be


imposed on the pier and estimate the nominal resistance of the structural
component or ground. Pier components shall satisfy the following
equation for each limit state:
i i Qi Rn = Rf
where:
i = Factors to account for ductility, redundancy and operational
importance
i = Load factor (dim)
Qi = Force effect or stress (kip)
= Resistance factor (dim)
Rn = Nominal resistance (kip)
Rf = Factored resistance (kip)
5.5.1.4

Strength Limit State Evaluations

The above equation should be used to evaluate piers and pier foundations
at the strength limit states for:

Bearing resistance failure

Lateral sliding

Excessive loss of base contact (eccentricity)

Pile group failure

Structural failure

The factored resistance, Rf, calculated for each mode of failure, is to be


calculated using the appropriate resistance factors for bearing resistance,
sliding, eccentricity, axial pile resistance and structural resistance.
The Designer should consider the consequences of changes in the pier
foundation conditions from scour due to the design flood event using
appropriate strength limit state resistance factors. Debris loading during
flood events should be accounted for in water pressure loads by assuming
a 25% increase in the exposed surface area of the pier.

March 2014

5-58

CHAPTER 5 - SUBSTRUCTURES

The investigation of piers at the strength limit states for structural failure
should be in accordance with LRFD Article 5.7 and carry all flexure and
axial loads anticipated. Appropriate consideration should be given to the
effects of slenderness on both aesthetics and load-carrying capacity.
For piers founded on piles, the shear on the critical section should be
investigated at the strength limit state in accordance with AASHTO LRFD
Section 5.13.3.6.
5.5.1.5

Service Limit State Evaluations

Piers should be investigated at the service limit state for:

Settlement

Lateral displacement

Overall slope stability

Foundation stability, settlement and horizontal movement at the


design flood for scour

A resistance factor, , of 1.0 is used to assess pier design at the service


limit state. The overall global stability of the foundation should be
investigated at the Service Load Combination with a resistance factor, ,
of 0.65.
Tolerable vertical and lateral displacement criteria for piers shall be
developed based on the function and type of pier, anticipated service life,
and consequences of unacceptable movements of the pier and effect on
the superstructure and bearings.
5.5.1.6

Extreme Event Limit State Evaluations

Extreme event limit state design checks for piers should include:

March 2014

Bearing resistance

Eccentricity

Sliding

Overall stability

Pile group failure

Structural failure

5-59

CHAPTER 5 - SUBSTRUCTURES

A resistance factor, , of 1.0 is used in the load and resistance equation in


Section 5.4.1.3 to assess pier design at the extreme limit state.
Resistance factors for extreme event limit states shall be taken as 1.0.
For the extreme event limit state, the Designer should consider scour due
to the check flood event and should determine that there is adequate
foundation resistance to support all applicable unfactored loads with a
resistance factor of 1.0 or less. Debris loading during flood events should
be accounted for in water pressure loads by a 25% increase in the
exposed surface area of the pier.
5.5.1.7

Structural Design

The structural design of piers shall be in accordance with the provisions of


LRFD Sections 5, 6, 7, and 8, as appropriate.
The investigation of piers at the strength limit states for structural failure
should be in accordance with LRFD 5.7 and carry all flexure and axial
loads anticipated. Appropriate consideration should be given to the
effects of slenderness on both aesthetics and load-carrying capacity.
For piers founded on piles, the shear on the critical section should be
investigated at the strength limit state in accordance with AASHTO LRFD
5.13.3.6.
5.5.1.8

Structural Design of Columns

The primary checks for a pier shaft or column structural design consist of:

March 2014

Determine maximum moments and shears in the shaft/column

Check limits for reinforcement (LRFD 5.7.4.2)

Calculate the factored axial resistance (LRFD 5.7.4.4)

Check slenderness provisions for compression members


(5.7.4.3)

Calculate the moment magnification factors (LRFD 4.5.3.2.2b)


Develop shaft or column interaction curve

Check biaxial flexure provisions for non-circular members


(LRFD 5.7.4.5)

Determine transverse reinforcement for compression members


(LRFD 5.10.6 or 5.7.4.6)

5-60

CHAPTER 5 - SUBSTRUCTURES

5.5.1.9

Geotechnical Design of Pier Foundations

A. Spread Footings
In using spread footings for foundation support for mass piers, either on
soil or bedrock, the design should be in accordance with the AASHTO
LRFD and Section 5.3 Spread Footings.
B. Deep Foundations
Deep foundations for mass piers may consist of piles or drilled shafts.
Piles may consist of H- or pipe pile steel sections, or precast concrete.
In founding a mass pier on a deep foundation, design should be in
accordance with the AASHTO LRFD, and BDG Sections 5.7 Piles and
5.8 Drilled Shafts. In designing deep foundation elements for a mass
pier with an unsupported length, a complete analysis of the foundation
should be performed using actual loading and soil conditions.
For strength and extreme limit state analyses, maximum factored axial
pile loads and stresses should be computed using 3-D pile group
analysis software, such as FB-Multipier.
For service limit state design of deep foundation, a complete deflection
analysis of a driven pile foundations should be performed using LPILE
or FB-Multipier software.
C. Scour
For scour protection of mass piers in water channels, the following
treatments should be considered: 1) the use of a deep seal placed
minimum of 2 feet below the scour depth determined for the check flood
for scour, or 2) designing the deep foundation elements for an
unsupported length. The unsupported pile length should be the vertical
distance from the bottom of the seal to the scour depth determined for
the check flood event. Piles should achieve axial capacity and lateral
capacity/fixity below the scour depth determined for the design flood
event.
5.5.1.10 Pier Protection
A. Collision Forces
Where the possibility of collision exists from vehicular, railroad, or water
traffic, an appropriate risk analysis should be made to determine the
degree of impact resistance to be provided and/or the appropriate
protection system.

March 2014

5-61

CHAPTER 5 - SUBSTRUCTURES

Unless the department determines that site conditions indicate


otherwise, or unless protected by collision walls as specified in
paragraph B. below, piers located within a distance of 30 feet to the
edge of roadway or within a distance of 50 feet to the centerline of a
railway track shall be designed for an equivalent static force of 400 kips,
which is assumed to act in any direction in a horizontal plane, normal to
the wall, at a distance of 4 feet above the ground.
B. Collision Walls
The provisions of the paragraph above need not be considered for piers
or abutments protected by an:
o An embankment
o A structurally independent crashworthy ground mounted 54 inch
high barrier, located within10 feet of the pier, or
o A 42 inch high barrier located at more than 10 feet from the pier
C. Vessel Collision
All bridge components in navigable waterway crossings where vessel
collision is anticipated shall be designed for a specified degree of vessel
impact damage in accordance with LRFD 3.14, or adequately protected
by dolphins, fender systems or other sacrificial devices.
D. Scour
The majority of bridge failures in the United States are the result of
scour. The added cost of making a bridge less vulnerable to scour is
small in comparison to the total cost of a bridge failure.
LRFD 3.7.5 requires that scour at bridge piers be investigated for two
conditions:
o For the design flood for scour, the streambed material above the
total scour line shall be assumed to have been removed. The
design flood storm event shall be the more severe of the 100-year
event or from an overtopping flood of lesser recurrence interval.
The strength and service limit states apply.
o For the check flood for scour, the stability of pier foundations shall
be investigated for scour conditions resulting from a designated
flood event, not to exceed the 500-year event or from an
overtopping flood of lesser recurrence. The extreme event limit
state shall apply. Reserve capacity beyond that required for
stability under this condition is not necessary. The exception is
March 2014

5-62

CHAPTER 5 - SUBSTRUCTURES

spread footings on soil or erodible rock, which shall be located to


that the bottom of the footing is below the scour depth determined
for the check flood for scour.
Refer to Section 2.3.11 Scour for additional guidance.
E. Facing
Where appropriate, the nose of the pier should be designed to effectively
break up or deflect floating ice or debris. Pier life can be extended by
facing the nose with steel plate/angle or by facing the pier with granite.
5.5.2 Pile Bent Piers
Pile bent piers are significantly less expensive than mass concrete piers and
provide environmental advantages by eliminating cofferdam work and its
associated impacts. Pile bents should be used wherever possible based upon
the criteria below.
5.5.2.1

Pile Bent Use Criteria

Pile bent piers should not be used in the following locations:

In rivers known for severe ice conditions - Allagash,


Androscoggin, Aroostook, Kennebec, Penobscot, St. Croix, and
St. John

Other locations with severe ice conditions

In shipping channels

Where the pier is not aligned with the design flow

Pile bent piers should be considered for structures in the following


locations:

March 2014

In tidal rivers

In environmentally sensitive areas

For grade-separated structures

Within the headwater or tailwater of dams or lakes, except when


ice has been known to form predominantly on one side of the
pier with an open channel in the adjacent span, resulting in
static ice forces on all piles.

5-63

CHAPTER 5 - SUBSTRUCTURES

The following issues affect the design of pile bent piers and must also be
considered when evaluating the appropriateness of this system.

Pile length - The pile length is a function of the depth to


bedrock, loading conditions, the type of overburden material, the
depth of scour, degree of pile fixity and restraint, and the pile
bracing.

Pile loads - The following issues affect pile loads:


1. Application location and magnitude of ice load
2. Skew - Longitudinal superstructure forces are transmitted
into the longitudinal pier axis and increase with greater skew
angles.
3. Bridge width - Pier cap shrinkage forces increase with
increasing bridge width.
4. Span length - Dead and live load axial forces are dependent
upon span length.
5. Seismic forces.

An additional issue to be considered when evaluating the appropriateness


of pipe pile pier bents is corrosion. Special consideration should be given
to corrosion and abrasion of steel pile bent piers to ensure a minimum 75
year structure life is achieved. This is of particular concern in locations
where there is insufficient water to install cathodic protection in
accordance with Section 5.5.2.6, and in locations where debris or
sediment loads may abrade pile protective coatings. In these locations
the design should consider additional protection such as encased H-piles
with sacrificial steel pipe pile or sacrificial fiber reinforced polymer (FRP)
composite pipe pile casings.
5.5.2.2

Loads and Load Combinations

Pile bent piers should be designed in accordance with AASHTO LRFD.


Structural analysis and design of reinforced concrete should include pile
bent cap flexure and shear checks, pile structural resistance and buckling
and lateral stability of piles. Geotechnical design checks should include
strength limit state checks and service limit state checks such as global
stability, horizontal bent displacement and pile settlement.
Where applicable, consideration should be given to other loading
conditions, including seismic forces resulting from earthquake loading and
debris lodged against pier, as outlined in 5.5.1.2 Load Combinations.

March 2014

5-64

CHAPTER 5 - SUBSTRUCTURES

Pile bent piers should be designed and proportioned to resist all applicable
load combinations specified for mass piers in 5.5.1.2 Load Combinations
and Load Factors, and as outlined in Chapter 3 Loads and LRFD Articles
3.4.1, 11.5 and 11.7.
A. Live Loads
Vehicular live loads must be located within the design lanes on the
superstructure such that maximum forces occur in the pile cap and piles.
Impact should be applied to pier caps and that the portion of the piles
that are acting as columns, defined as the vertical distance from the pile
cap to the point of fixity below grade. Impact should be applied at or
above Q1.1.
B. Ice Loads
For the Extreme Event II load combination, unfactored ice loads should
be placed at the Q50 stage elevation and checked at a lower elevation
that will cause maximum moment in the nose pile, provided the elevation
is at or above Q1.1. The ice thickness of past practice should be
increased by 1.0 foot.
Transverse ice loads should be applied to only the nose pile when ice is
directly applied to the nose pile, or be uniformly distributed over the cap
when ice is applied to the cap.
A modified Strength Limit State analysis should also be performed with
factored ice loads following the criteria specified in 3.9 Ice Loads, with
appropriate strength limit state resistance factors for the pier component
being analyzed.
C. Water Loads
Stream pressure should be reduced when the ice elevation is lowered to
check maximum moment in the nose pile.
Stream pressure should be applied to each pile in the bent, using an
appropriate stream flow velocity.
D. Wind Loads
Longitudinal components of wind on superstructure and wind on live load
should be distributed to the abutments when structure fixity is at the
abutments.

March 2014

5-65

CHAPTER 5 - SUBSTRUCTURES

E. Seismic Loads
Seismic loads transverse to the bridge should be shared between all
substructure units based upon their stiffness.
Longitudinal seismic loads should be distributed to the abutments where
there is at least one fixed abutment with no forces applied to the pier.
F. Shrinkage and Temperature Forces
Shrinkage and temperature forces affect pile bents in two ways:
o Pile cap shrinkage and temperature actions are applied to the
longitudinal axis of the pier.
o Thermal forces are induced by the superstructure are applied
along both the transverse and longitudinal pier axes, with the
magnitude dependent upon the skew angle.
Two-span integral abutment bridges will have no associated thermal
forces applied, as the forces are assumed to be balanced at the pier.
The Structural Designer may want to include thermal forces for two-span
integral abutment bridges on steep grades, assuming that the bridge will
expand and contract downhill.
For non-integral abutment bridges, thermal forces induced by the
superstructure bending the pile bents must be considered in the design
of the fixed abutment.
G. Braking Forces
If the structure is fixed at an abutment, the longitudinal braking forces will
have no effect on the pier, as the forces are assumed to be distributed to
the abutments.
H. Friction Forces
Friction forces resulting from all longitudinal superstructure forces should
be applied to pile bents with expansion bearings.
I. Collision Loads
Where the possibility of collision exists from vehicular, railroad, or water
traffic, an appropriate risk analysis should be made to determine the
degree of impact resistance to be provided and/or the appropriate
protection system.

March 2014

5-66

CHAPTER 5 - SUBSTRUCTURES

5.5.2.3

Pile Cap Design

Pile bent cap design should consider the following design features:

Piles should be embedded at least 12 inches

Pile clearance with 6 inches of concrete cover

Tolerance on pile installation misalignments > or = 2 inches

Consider concrete pile anchorage

Pile spacing should be at least 30 inches or 2.5 times the pile


diameter

5.5.2.4

Pile Type Selection Criteria

Concrete filled pipe piles, precast concrete piles, combination H-piles


encased with pipe piles filled with concrete, and drilled shafts may be
considered for pile bent piers under the following conditions:
A. Shallow overburden depth (embedment less than or equal to the fixity
depth)
o Footing-encased pipe or precast concrete piles
o Rock-socketed pipe piles
o Rock-socketed H-piles, with pipe pile encasement to top of
bedrock
o Rock-anchored/doweled pipe piles (Note: AASHTO LRFD is
absent of discussion on the use of rock-anchor pipe piles. The
use of rock-anchored pipe piles should be considered only when
the preceding alternatives are found not feasible. Rock anchors
or dowels should have double corrosion protection.)
o Rock-socketed drilled shafts
B. Intermediate overburden depth (embedment greater than depth to
fixity and less than 3 times fixity depth)
o Pipe piles filled with concrete and a reinforcing cage (The
reinforcing cage may be eliminated with the approval of the
Engineer of Design.)
o Precast concrete piles

March 2014

5-67

CHAPTER 5 - SUBSTRUCTURES

o Drilled shafts
C. Deep overburden depth (embedment greater than 3 times fixity
depth)
o Pipe piles filled with concrete and a reinforcing cage (The
reinforcing cage may be removed with the approval of the
Engineer of Design.)
o H-piles with pipe pile encasement to pile fixity depth
o Precast concrete piles
o Drilled shafts
The choice of steel versus concrete piling in intermediate and deep
applications should be determined by a cost analysis. Issues include the
relative costs of H-piles to precast concrete piles or pipe piles,
encasement and the relationship between the exposed length (including
the scour depth), the depth to fixity, and the total depth to bearing.
D. Pier Bent Pile Alternatives
Because of ongoing corrosion and durability issues with steel pipe piles,
Geotechnical Engineers and Designers should routinely examine the
feasibility and practically of other pier bent pile-types, namely:
o precast concrete piles
o drilled shaft pier bents
o encased H-piles with a sacrificial steel pipe pile or a sacrificial
fiber reinforced polymer (FRP) composite pipe pile casing
5.5.2.5

Pile Protection

A. Encased H-Piles
Steel H-piles should not be used for piers without full encasement
protection. The encasement usually is a steel pipe pile filled with
concrete. H-piles should be protected by a minimum of 3 inch clear
encasement from the pier cap to a minimum of 10 feet below streambed
or 2 feet below the total scour depth. Due to the significant additional
load section provided by the composite steel and concrete section, the
pipe pile should be used for strength. If the pipe pile is used for strength,
it should extend to the point of fixity below streambed.

March 2014

5-68

CHAPTER 5 - SUBSTRUCTURES

The pipe pile should be protected and designed as detailed in Paragraph


B. Pipe Piles, below.
B. Pipe Piles
Pipe piles bents in fresh water environments should be hot-dipped
galvanized with UV-resistant epoxy top coat. Pipe pile bents in brackish
or salt water should be coated with fusion bonded epoxy paint with a
coat thickness of 18-20 mils. This is an increase in the previous
standard of 12 mils.
Fusion-bonded epoxy coatings or galvanized surfaces should be applied
to a minimum of 10 feet below streambed or 2 feet below the total scour
depth.
Cathodic protection (aluminum anodes) should always be used in
addition to the protective coatings in salt and fresh water environments
as long as there is sufficient water to submerge the anodes at low water.
Refer to 5.5.2.6 Pipe Pile Coatings and Cathodic Protection for detailed
recommendations.
C. Precast/Prestressed Concrete Piles
Concrete cover for rebar should be a minimum of 2 inches for fresh
water locations and 3 inches for salt water locations.
5.5.2.6

Pipe Pile Coatings and Cathodic Protection

A. Standardized Anodes
Pipe pile pier bents and cargo/ferry piers should specify a standard
anode ingot length, composition (aluminum alloy plus minor constituents)
and weight.
The standard should be a 34-lb, aluminum alloy anode, approximately 3
feet long. Larger, heavier anodes are not easy to handle and should be
avoided unless the bent has a lot of uncoated steel or the project is a
significant sheet pile structure where there is a greater chance for
exposed steel. On large piles with long exposed lengths (deep water),
consideration should be given to installing more than one anode rather
than using a heavier anode.
There are a lot of variables in the rate of corrosion between sites, and it
may happen that the standard anode may not be suitable for all sites.
Larger anodes may be necessary for more aggressive environments
(brackish and saltwater). Specifying a heavier anode may be required.

March 2014

5-69

CHAPTER 5 - SUBSTRUCTURES

B. Anode Location
The top of the 34-lb, 3-foot long anode should be 3 feet below Low
Water, so that it is always submerged. This implies the water channel
needs to have at least 6 feet of water at Low Water.
Anodes should be installed on the more protected side of the pile: on
the underside on battered piles, on the downstream side on plumb piles
in rivers, and on the more protected side (if there is one) on plumb piles
in tidal crossings. If possible, show the location of the anodes on the
Plan drawings, so there is no debate in the field about what constitutes
the more protected side.
C. Shallow Water Situations
If there is not enough depth of water to submerge the anodes at all
times, the anodes are not as effective in protecting the pile segment
above the waterline.
Where the water is shallow and there is no submerged portion or a
limited submerged portion of pile for anodes, Designers should consider:
o specify a non-standard, shorter ingot if that permits installation on
a pile in shallow water
o fusion-bonded epoxy treatment over hot-dipped galvanized piles
o encasing H-piles with a sacrificial steel pipe pile or a sacrificial
FRP composite pipe pile casing
D. Anode Attachment
Plans should specify a 2-inch clearance between the anode and the pile.
This allows Bridge Inspectors to get a clear view of the anode, and the
pile surface is more inspectable and the anodes easier to replace.
Attachment hardware consisting of a 3-inch long, -inch diameter
threaded stud, with double nuts, is recommended. The studs should be
installed in a manner that ensures the best steel to steel connection and
the best electrical connection. The weld area shall be ground to bare
metal for this purpose. Only after the stud and anode are attached, shall
the weld at the base on the stud be covered with curable polyamide
epoxy coating.
E. Brackish and Saltwater Environments
Steel pile bents in brackish or salt water should not be hot-dipped
galvanized with UV-resistant epoxy top coat. These pile bents should be
coated with fusion bonded epoxy coating with a thickness of 18-20 mil.
March 2014

5-70

CHAPTER 5 - SUBSTRUCTURES

Cathodic protection aluminum anodes should always be installed on pile


pier bents in salt water when there is enough water.
F. Freshwater Environments
Steel pile bents in freshwater should be hot-dip galvanized with a UVresistant topcoat system. The UV-resistant topcoat tends to fade where
the upper part of the pile gets direct sunlight and reflected light from the
water surface. Considerations should be given to topcoating with
fluorocarbon paint, which is more UV resistant.
Cathodic protection aluminum anodes should be installed on pile pier
bents in fresh water, with the exception of river crossings with very
shallow water.
G. Coating Repairs
Pile coating touch-up per the manufacturers recommendations is
considered the best practice for dealing with damaged pile sections.
The touch-up material on some jobs (in particular, Alna-Newcastle)
seems to be performing well. The Bridge Program should determine the
best touch-up method and specify it not just specify touch-up per
Manufacturers recommendations.
H. Pipe Pile Material
Steel pipe piles should be ASTM 252 and have straight butt-welded
seams or be seamless. Spiral seams are not recommended because
the magnitude of welded surfaces which are vulnerable to thin coatings,
ice abrasion, and bumping during construction all of which lead to
damage in the coating. Welds should be ground down and blended
smoothly with the pile material. The number of field and mill splices
should be limited.
I. Damage during Construction
Specifications should include requirements for Contractors to line driving
templates with fire hoses, carpets, etc., to prevent the scraping off of the
coatings during pile driving. Contractors should be required to repair or
replace any protective mats that fall off during driving, prior to
commencing driving any more pile.
5.5.2.7

Additional Pile Bent Pier Design Criteria

Pile bents should consist of a concrete pile cap supported by a single row
of piles, multiple rows of piles, or a braced group of piles.

March 2014

5-71

CHAPTER 5 - SUBSTRUCTURES

A. Pile Design
Pile design should investigate resistance to axial loads, combined axial
and bending, and buckling failure of the exposed pile lengths. Guidance
for computing the unsupported pile length is provided in Section B,
below. Stability of the pile bent pier under combined axial and lateral
loads should be investigated with a dedicated soil-structure interaction
analysis, using FB-Pier software.
B. Pile Length
The unsupported length, Lus, is defined by the following:

L us K (L u L e )
where:
K=

Effective Length Factor. Refer to LRFD Article 4.6.2.5 and


Table C4.6.2.5-1.
Lu = Exposed pile length above ground.
Le = Effective pile length from ground surface to the point of
assumed fixity below ground, including scour effects. Refer to
Figure 5-4 and Figure 5-5.
The depth to fixity shown in Figure 5-4 and Figure 5-5 was determined
using the Davisson and Robinson procedure provided in LRFD Article
10.7.3.13.4 and assumes no lateral loading on the pile. Where piles used
for pile bent piers are subjected to lateral loading or where the embedment
length is less than 3Le, a detailed analysis by the Designer using actual
loading and soil conditions is required.

March 2014

5-72

CHAPTER 5 - SUBSTRUCTURES

20

Effective Pile Length, Le, from Ground Surface


to Depth of Fixity (ft)

18
loose
medium dense
dense

16
14
12
10
8
6
4
2
0
10

100

1000

Moment of Inertia (in4)

Figure 5-4

10000

Effective Pile Length for Piles in Sand

From Ground Surface to Depth of Fixity


Axially Loaded

March 2014

5-73

CHAPTER 5 - SUBSTRUCTURES

24

Effective Pile Length, Le, from Ground Surface


to Depth of Fixity (ft)

22
20

Su = 375 psf
Su = 750 psf
Su = 1125 psf

18
16
14
12
10
8
6
4
2
0
10

100

1000

Moment of Inertia (in 4)

10000

Figure 5-5 Effective Pile Length for Piles in Clay


From Ground Surface to Depth of Fixity
Axially Loaded
C. Nose Pile Batter
Where possible, the nose pile should be battered a minimum of 15 to
take advantage of the allowance for ice load reduction due to nose
inclination (refer to LRFD Article 3.9.2.2). When ice is applied to the pier
cap or within 5 feet of the pier cap, no reduction should be taken.
D. Design Section
Encased H-piles and concrete-filled pipe piles should be designed
assuming contribution from the concrete and a portion of the steel pipe
pile shell, allowing for a minimum of 0.15 inch of sacrificial shell
March 2014

5-74

CHAPTER 5 - SUBSTRUCTURES

corrosion, based on a design corrosion rate of 0.05 mm per year. The


pipe pile shell must have a minimum thickness of 1/2" to allow for proper
driving of the pile and to resist corrosion.
MaineDOT Section 711.01 specifies ASTM 252 for Welded and
Seamless Steel Pipe Piles. Designers should consider that ASTM 252
permits under-fabrication of the wall thicknesses up to 12.5% of the
specified nominal wall thickness. Example: If the design calls for 5/8inch wall, the design section should be reduced by a minimum 1/8-inch
for sacrificial steel shell corrosion and an additional 1/16-inch to account
for permissible fabrication variation.
5.6

Retaining Walls

5.6.1 General
Retaining walls typically used by the Bridge Program are gravity walls,
cantilever-type walls, mechanically stabilized earth (MSE) walls, prefabricated
proprietary walls and soil nail walls, each of which is discussed in detail in the
following sections. The selection of the appropriate retaining wall should be
based on an assessment of the magnitude and direction of loading, depth to
suitable foundation support, potential for earthquake loading, presence of
deleterious factors, proximity of physical constraints, wall site cross-section
geometry, tolerable and differential settlements, facing appearance, and ease
and cost of construction. A feasibility study should address which wall is most
suited to the site and is simplest to construct. The study should address the
approximate scope of the design for the most feasible walls, and provide cost
comparison between alternatives.
5.6.1.1

Retaining Wall Type Selection

Due to construction techniques and base width requirements, some wall


types are best suited for cut sections whereas others are best suited for fill
situations. The key considerations in deciding which wall is feasible are
the amount of excavation or shoring required and the overall wall height.
The site geometric constraints must be well-defined to determine these
elements.
A. Walls in Cut Sections
Anchored walls and soil nail walls, which have soil reinforcements drilled
into the in-situ soil/bedrock, and cantilever sheet pile walls, are generally
used in cut situations. These walls are typically constructed from the top
down.

March 2014

5-75

CHAPTER 5 - SUBSTRUCTURES

B. Walls in Fill Sections


MSE walls are constructed by placing soil reinforcement between the
layers of fill from the bottom up and are therefore best suited to fill
situations. Additionally, the base width of MSE walls is typically on the
order of 70% of the wall height, which would require considerable
excavation in a cut section, making the use of this wall uneconomical.
C. Walls in Cut or Fill Sections
Gravity, cantilever-type, and prefabricated proprietary walls are
freestanding structural systems built from the bottom up that do not rely
on soil reinforcement techniques to provide stability. These types of
walls have a narrower base width than MSE structures (on the order of
50% of the wall height) making this type of wall feasible in fill situations
as well as many cut situations.
5.6.1.2

Service Life

Retaining walls should be designed for a service life based on


consideration of the potential long-term effects of material deterioration,
seepage, stray currents, and other potentially deleterious environmental
factors on each of the material components comprising the wall. For most
applications, permanent retaining walls should be designed for a minimum
service life of 75 years. Retaining walls for temporary applications are
typically designed for a service life of 36 months or less. Greater level of
safety and/or longer service life (i.e., 100 years) may be appropriate for
walls that support bridge abutments, for which the consequences of poor
performance or failure would be severe.
The quality of in-service performance is an important consideration in the
design of permanent retaining walls. Permanent walls should be designed
to retain an aesthetically pleasing appearance, and be essentially
maintenance free throughout their design service life.
5.6.1.3

Design Loads

Retaining walls should be designed in accordance with AASHTO LRFD


Structural analyses, the design of reinforced concrete and geotechnical
analyses of retaining walls will be computed using LRFD procedures using
factored loads and factored resistances. The geotechnical design of
conventional retaining walls typically follows the LRFD approach for the
design of abutments on spread footings, presented in 5.3 Spread Footings
and 5.4 Abutments. Where a wall is supported with piles or dilled shafts,
the design will follow LRFD and 5.4.1.12 Abutments Supported on Pile
Foundations and 5.7 Piles, as appropriate. Loads should be determined
in accordance with AASHTO LRFD and as outlined in Chapter 3 and
March 2014

5-76

CHAPTER 5 - SUBSTRUCTURES

5.4.1.2 Load Combinations and Load Factors. The following load


conditions should be considered when applicable:

Lateral earth pressure

Weight of soil above the footing or within the wall system

Self-weight of the wall

Lateral loads due to live load impact on the parapets

Surcharge loads, due to live load

Surcharge load caused by earth, point, line or strip loads on the


upper surface

Railroad loading

Hydrostatic pressure (if no drainage is provided)

Earth pressure due to compaction should be considered when static or


dynamic compaction is used within a distance of one-half of the wall
height. These loads will only apply to during construction phase; therefore
a load factor of 1.0 is appropriate.
Walls that can tolerate little or no movement, or are restrained, should be
designed for at-rest (Ko) earth pressure with a maximum load factor for atrest earth pressure, EH, of 1.35.
5.6.1.4

Limit States

Retaining walls should be designed to resist all applicable load


combinations specified in LRFD 3.4.1 and 11.5.5.
Strength limit state checks of walls should assess external failure
mechanisms:

Sliding

Eccentricity

Bearing Resistance

Structural Capacity

Service limit state check should assess overall stability, wall settlement
and lateral displacement.
Walls should be evaluated for each of the applicable limit states:
March 2014

5-77

CHAPTER 5 - SUBSTRUCTURES

Strength I-construction. Strength Limit State I which models the


basic load combination related to construction loads. Load
factors for the dead load of other components shall not be less
than 1.25. Live load surcharge is included to account for
construction equipment live loading; a construction load factor of
not less than 1.5 should be assumed.

Strength I-a: Strength Limit State I, which models the basic load
combination related to normal vehicle live load surcharge, dead
load plus earth pressure, finished grade, including any point or
strip loads on the wall backfill Minimum vertical permanent load
factors and maximum horizontal load factors are selected to
produce extreme force effects for wall sliding and eccentricity,
and structural design of the wall stem.

Strength 1-b: Strength Limit State 1 as described above, except


maximum vertical permanent load factors, including earth loads,
are selected to produce an extreme force effect for bearing
capacity analyses.

Service I: Service Limit State I Load combination relating to


normal operational use of the wall with all loads taken at their
unfactored values.

Wall foundations subject to scour should be designed at the strength and


service limit states so that there is adequate foundation resistance, in
conjunction with the depth of scour from the design flood, using
appropriate strength and service limit state resistance factors.
The consequences of changes in wall foundation conditions due to scour
from the check flood for scour should be assessed at the extreme event
limit state with resistance factors of 1.0.
Where retaining walls are to be designed to resist earthquake forces,
collisions by roadway or railway vehicles, or vessel collision, the structures
should be evaluated for the following additional limit states:

Extreme Event I Load combination including earthquake


forces

Extreme Event II Load combination relating to collision by


vehicles, railways or vessels.

Each load for each limit state above is modified by the prescribed load
factor, . Certain permanent loads, including earth loads, should be
factored using the load factors p. Load factors should be selected to
produce the total extreme factored force effect. Applicable load factors,

March 2014

5-78

CHAPTER 5 - SUBSTRUCTURES

load combinations and the analyses for which they will govern, are
provided in Table 5-6.
Table 5-6 Typical Load Groups and Load Factors
Load
Group

DC

EV

LSV

Strength 0.90
I-a

1.0

1.75 1.75

1.5

1.35

1.5

Strength 1.25 1.35 1.75 1.75


I-b

1.5

1.35

1.5

Service
I

1.0

1.0

1.0

5.6.1.5

1.0

1.0

1.0

LSH

1.0

EH
EH ES
(active
(at&
rest)
passive)

Typical
Geotechnical
Analysis
Governed
Sliding
Eccentricity
(overturning)
Structural
design of wall
stem
Anchor
pullout
Bearing
Capacity
Structural
design of the
wall footing
Settlement
Lateral
displacement
Global
stability

Strength Limit State

A. Bearing resistance
The check for bearing resistance for wall spread footings on soil or rock
is identical to the requirements for abutments described in 5.3.5 Bearing
Resistance. Wall foundations subject to scour should be designed so
that the nominal bearing resistance, in conjunction with the depth of
scour determined for the check flood for scour, provides adequate
resistance to support the unfactored Strength Limit State Loads with a
resistance factor of 1.0.
B. Eccentricity
The overturning calculation used in ASD is replaced with the eccentricity
check. Eccentricity of loading on walls founded on spread footings is
identical to the requirements for abutments, should be calculated for
each load group and checked to meet the following criteria:

March 2014

5-79

CHAPTER 5 - SUBSTRUCTURES

o E < B/3 for foundations on soil


o E < 0.45B for foundations or bedrock
C. Sliding
Sliding calculations for walls on spread footings are identical to the
requirements for abutments described in 5.3.8 Sliding. Passive
pressures in front of the wall should be neglected. To maximize the
effect of the live load surcharge, the horizontal component of the live
load surcharge should be included, whereas the vertical component over
the heel or base should be neglected.
D. Pile Resistance
The design of walls founded on deep foundations is similar to the design
requirements described in 5.4.1.12 Abutments Supported on Pile
Foundations and 5.7 Piles.
E. Overall Stability
The overall global stability of retaining walls should always be checked
at Service I load combination with a resistance factor, , of 0.65.
5.6.1.6

Service Limit State Checks

Service limit state wall settlement should be checked with the following
performance limits in mind:

March 2014

Total settlement can be estimated using the procedures and


criteria described in 5.3.6 Settlement. The tolerable total
settlement criterion generally considers its effect on
serviceability.

Settlement may be critical where the wall interacts with other


structures, e.g. at the approach to a pile supported abutment.

Distortion, i.e., the ratio of horizontal movement to vertical


movement should be less that 1/500.

Lateral deformations will usually take place during construction


and be affected by wall batter, compaction effort and
construction equipment next to the wall.

Global stability.

5-80

CHAPTER 5 - SUBSTRUCTURES

5.6.1.7

Design Considerations

All retaining walls should be designed with consideration of frost protection


(Section 5.2.1), scour protection (Section 2.3.11), bearing resistance
(Section 5.3.5), settlement (Section 5.3.6), stability (Section 5.3.7),
drainage considerations (Section 5.3.11), and seismic considerations
(Section 5.2.5), as appropriate.
All retaining walls require a subsurface investigation of the underlying soil
or bedrock that will support the structure or tie-back elements. Minimum
requirements for number, spacing and depth of exploratory borings are
provided in Section 2.10 Subsurface Exploration Programs.
5.6.1.8

Aesthetics

Retaining walls should have a pleasing appearance that is compatible with


the surrounding terrain and other structures in the vicinity. Aesthetic
requirements include consideration of the wall face material, the top
profile, the terminals, and the surface finish (texture, color, and pattern).
Where appropriate, provide planting areas and irrigation conduits. In
higher walls, variation in treatment is recommended for a pleasing
appearance. High, continuous walls are generally not desirable from an
aesthetic standpoint. Consider stepping high or long retaining walls in
areas of high visibility.
5.6.2 Gravity Retaining Walls
Gravity retaining walls are generally trapezoidal in section and derive their
capacity to resist lateral soil loads through a combination of self- weight and
sliding resistance. Gravity walls can be subdivided into rigid gravity walls,
which will be discussed in this section, MSE walls discussed in Section
5.6.5.4, and prefabricated proprietary walls discussed in Section 5.6.5.
5.6.2.1

Design Section

Gravity wingwalls should have a thickness at the top of 1-6 in a direction


normal to the front neat line. Batters on the front and back faces of
wingwalls should be related to the vertical plane, which is normal to the
front neat line. The front neat line is a horizontal line, which is the
intersection of the top of footing elevation and the front face of the wall. If
there is no footing, a working elevation should be used. Gravity walls of
any length should be constructed to work integrally with abutments.

March 2014

5-81

CHAPTER 5 - SUBSTRUCTURES

5.6.2.2

Earth Loads

Rigid gravity walls should be designed as unrestrained, which means that


they are free to rotate at the top in an active state of earth pressure. An
active earth pressure coefficient, Ka, should be calculated using Coulomb
Theory as described in Section 3.6.5.1.
5.6.3 Gravity Cantilever-type Retaining Walls
This section discusses gravity, cantilever-type retaining walls. This type of
wall is differentiated from a non-gravity cantilever retaining wall by relying on
the bending action of the wall stem, in addition to self-weight, to resist lateral
earth pressures. The footing contributes to the wall stability in overturning and
sliding. Non-gravity cantilever retaining walls (i.e., sheet pile walls) are
discussed in Section 5.6.4.
5.6.3.1

Design Section Gravity Cantilever Retaining Walls

Cantilever walls should have the following limits for wall thicknesses
(heights are measured from top of the wall footing):

1-3 minimum thickness for walls up to 6 feet high at the highest


point.

1-6 minimum thickness for walls between 6 feet and 20 feet in


height at the highest point.

1-9 minimum thickness for walls over 20 feet in height at the


highest point.

Walls should be increased in thickness to accommodate


recessed architectural treatment, as necessary.

Wingwalls that are 15 feet or more in height at the ends may be designed
with butterfly wings, if economical to do so.
On wingwalls that are less than 15 feet in height at the ends, the footing
may be reduced in length if it is not required for structural or geotechnical
considerations. The wall should be detailed with the bottom of the wall at
the elevation of the top of the footing.
Tops of parapets should not have elevations above the adjacent curbs or
sidewalks.
Gravity cantilever wingwalls more than about 20 feet long should be
designed to work independently from the abutment, except that footings
should be integral. A vertical contraction or expansion joint with no shear

March 2014

5-82

CHAPTER 5 - SUBSTRUCTURES

key should be used near the corner between the abutment and the
wingwall. The front face of the wingwalls should be recessed 2 inches
back from the face of the wall on the abutment side of the contraction or
expansion joint.
Gravity cantilever type wingwalls that are less than about 20 feet long
should also be designed independently from the abutment; however, the
wingwall should be restrained at the corner through an integral connection
to the abutment. Soil pressure under the footing, sliding, and eccentricity
should be evaluated as discussed in Section 5.3 Spread Footings. The
restraining force at the corner is considered to be caused by at rest lateral
earth pressure, as a minimum, because of the wingwalls inability to
deflect at the corner. The corner should be designed to be restrained by
concrete beam action with horizontal reinforcing steel anchored into the
abutment section.
5.6.3.2

Earth Loads

For earth loads relative to cantilever walls refer to Section 3.6. Load
factors for earth loads and surcharge loads are provided in Table 5-4. In
the case of a long wall with a variable height, the wall should be divided
into more than one design section. The design section should be at the
highest third point of the wall. Refer to Figure 5-6 for further guidance.
Design Section

2/3 H

Figure 5-6 Retaining Wall Design Section


Gravity cantilever walls should be designed as unrestrained, which means
that they are free to rotate at the top in an active state of earth pressure.
An active earth pressure coefficient, Ka, should be as described in Section
3.6.4 and factored as specified in Table 5-6.

March 2014

5-83

CHAPTER 5 - SUBSTRUCTURES

5.6.4 Non-Gravity Cantilever and Anchored Retaining Walls


This section discusses non-gravity cantilever retaining walls. Non-gravity
cantilever retaining walls derive lateral resistance through embedment of
vertical wall elements, sometimes in combination with anchors or tie-backs.
These vertical elements may consist of sheet piles, soldier piles, caissons, or
drilled shafts. The vertical elements may form the entire wall face or they may
be spanned structurally using timber lagging or other materials to form the wall
face.
The design of cantilever and anchored walls include additional checks for the
geotechnical resistance of anchors in pullout, bearing resistance of vertical
elements, the passive resistance of vertical elements and the structural
capacity of anchors, vertical wall elements and wall facing. Resistance factors
specific to cantilever and anchored walls can be found in LRFD Table 11.5.6-1
5.6.4.1

Soil Nail Walls

Soil nail walls are technically anchored walls that employ a reinforced soil
mass serving as a gravity retaining structure. The reinforced soil mass of
a soil nail wall is created by drilling and grouting steel anchors into an insitu soil mass. The anchored soil mass is then covered with shotcrete.
The temporary shotcrete face is then covered with a permanent facing
system, typically cast-in-place concrete, precast concrete, or timber
lagging. Soil nail walls are suited to cut situations only.
Soil nail walls are relatively low cost and can be used in areas of restricted
overhead or lateral clearance. Soil nail walls are built from the top down
and are only suitable if the site soils have adequate stand-up time of 1 to
2 days in a 5 foot vertical cut. Soil nail walls are not applicable to sites
with bouldery soils, which could interfere with nail installation. This wall
type is not recommended in uniform or water bearing sands or where
there is a potential deep failure surface. Maximum wall heights of 30 feet
are allowed.
These walls can be designed by the Designer or specified as a designbuild item. The PS&E package should include the plan development
information discussed in Section 5.6.5.5. Special Provisions have been
developed for soil nail walls. Check with the Geotechnical Designer for
the current Special Provision.
5.6.5 Prefabricated Proprietary Walls
Prefabricated proprietary walls are any prefabricated wall system approved by
MaineDOT and produced by a manufacturer licensed by the wall vendor.
Prefabricated proprietary walls are typically designed by the vendor, but may
be designed by the Geotechnical Designer. In design, the vendor should
March 2014

5-84

CHAPTER 5 - SUBSTRUCTURES

consider external stability with respect to sliding and overturning (at every
module level) and internal stability with respect to pullout, as specified in LRFD
11.10 and Chapter 3, Loads. The Geotechnical Designer is required to verify
acceptable global stability of the wall using a resistance factor of 0.65 prior to
advertisement. The factored bearing resistance of the wall foundation soil or
bedrock must be shown on the plans.
5.6.5.1

Proprietary Retaining Walls

Retaining walls available for a given project include standard walls, where
the responsibility of the design is the Structural Designer, and proprietary
walls, which are designed by a wall manufacturer. There are MaineDOT
preapproved proprietary wall systems and non-approved proprietary wall
systems. Preapproved wall systems have been extensively reviewed by
MaineDOT and are listed on the MaineDOT Qualified Products List (QPL)
webpage for the particular wall type. MaineDOT has developed a review
process for the pre-approval of non-approved proprietary walls systems
(MaineDOT, 2010), available on the MaineDOT QPL website. Nonapproved proprietary walls must go through the pre-approval review
process prior to use of the wall system.
5.6.5.2

Prefabricated Concrete Modular Gravity Walls

Prefabricated concrete modular gravity (PCMG) walls covered under


Special Provision 635 should consist of either T-Wall as provided by a
licensed manufacturer of the Neel Company, Springfield, Virginia, or
DoubleWal as provided by a licensed manufacturer of DoubleWal
Corp., Plainville, Connecticut.
PCMG walls should be designed in accordance with Special Provision 635
and LRFD Article 11.11. In general, the design requirements are similar to
the requirements for conventional retaining walls and abutments, with the
exception of pullout resistance requirements and dedicated analyses at
each level of modular units.
PCMG walls should be considered on all projects where metal bin, gabion,
MSE, and cast-in-place walls are considered. PCMG walls should be
limited to a maximum height of 27.5 feet and a maximum batter of 1/6 (2
inches per foot). Refer to Section 5.6.5.5 PS&E for Project with
Proprietary Walls for plan development requirements.
Whenever possible, a battered wall will be used in preference to a vertical
wall. The use of a vertical wall design may be necessary where the wall is
located on a horizontal curve that may result in construction conflicts, or
where property costs or other right-of-way considerations dictate.

March 2014

5-85

CHAPTER 5 - SUBSTRUCTURES

PCMG walls should be designed with adequate embedment for frost


protection. Refer to Section 5.2.1 Frost for guidance.
PCMG walls should not be used in locations where there is scour
potential, unless suitable scour protection can be economically provided.
Refer to Section 2.3.11 Scour for guidance.
Where special drainage problems are encountered, such as seepage of
water in the excavated backslope, underdrain will be provided behind the
wall. Refer to Section 5.3.11 Drainage Considerations for further
guidance.
Where PCMG walls will come in contact with salt water, all rebar should
be epoxy coated and the concrete should be class LP. The appropriate
note from Appendix D Standard Notes Prefabricated Concrete Modular
Gravity Wall should be on the contract drawings.
Where PCMG walls are to be located in water, consideration should be
given to drainage behind the wall. As a minimum, the Designer should
consider a 12 inch thick layer of crushed stone extending vertically along
the inside wall face. Crushed stone should be separated from surrounding
soils with an erosion control geotextile. When drainage features are used
for PCMG walls, payment should be considered incidental.
PCMG walls may be considered to retain soil supporting bridge
substructures, with the exception of bridges over waterways. Their use is
subject to the approval of the Assistant Bridge Program Manager at the
PDR stage. These types of walls shall be designed for a service life of
100 years. The PCMG concrete shall contain a minimum of 5.5 gal/yd3 of
corrosion inhibitor and use corrosion resistant reinforcing. PCMG walls
which retaining abutments and are within 30 feet of the edge of a roadway
or 50 feet of the centerline of a railway track should be designed for
collision forces or protected with a crashworthy barrier (see 5.4.1.7.E).
Additional design criteria for abutments retained by PCMG walls are
similar to those for MSE walls described in 5.6.5.4.
Cofferdams required for PCMG wall construction should be considered
incidental to wall construction. The appropriate notes from Appendix D
Standard Notes Prefabricated Concrete Modular Gravity Wall should be
on the contract drawings.
PCMG walls are measured and paid for by the area of wall face, as
determined from the plan dimensions. The PCMG pay item includes
compensation for excavation, excavation support foundation material,
backfill material, and wall design. Consult Special Provision 635 for
current measurement and payment information.

March 2014

5-86

CHAPTER 5 - SUBSTRUCTURES

5.6.5.3

Precast Concrete Block Gravity Walls

Precast concrete block gravity walls consist of walls where precast


concrete units are stacked vertically, function either as a gravity retaining
wall or as a facing with geosynthetic-reinforced soil backfill, as covered in
Special Provision 635. The connection between adjacent courses of
modular blocks may be mechanical (cast knobs) or frictional. A
preference is for mechanical connections. These wall systems are
generally limited to a maximum height of 4.5 feet when the precast
concrete units function as a gravity wall without reinforced backfill and no
surcharge load is applied. When wall height is in excess of 4.5 feet or a
surcharge is applied, geosynthetic reinforcement may be added to the
modular blocks to create a geosynthetic-reinforced soil (GRS) wall.
Precast Concrete Block Gravity Walls without reinforced backfill should
meet the design requirements of LRFD 11.11. If the backfill is reinforced,
walls should meet the design requirements of LRFD 11.10 and BDG
5.6.5.4.B. Geosynthetic-Reinforced Soil Walls.
Blocks for modular block walls are made from wet cast concrete. Wall
systems comprised of dry cast concrete are susceptible to degradation
caused by freeze-thaw and are not an approved wall type. Precast
concrete block gravity walls are not permitted in waterways.
5.6.5.4

MSE Walls

A. MSE Walls with Steel Reinforcement


This type of MSE wall uses galvanized strips or mats of steel to reinforce
soil and create a reinforced soil block behind the wall face. The
reinforced soil mass acts as a unit and resists the lateral loads through
the dead weight of the reinforced mass. MSE walls are constructed from
the bottom up and are therefore best suited for fill situations.
With a few exceptions, the procedure for the design of MSE walls using
LRFD is identical to that followed using ASD. External stability
evaluations include bearing resistance, sliding, and eccentricity. Internal
stability calculations include pullout and rupture of reinforcements,
capacity of reinforcement connections to the wall face, and structural
capacity of the wall facing. MSE walls are typically designed by the wall
manufacturer for internal and external stability. All MSE walls should be
designed in accordance with:
1. LRFD Article 11.10

March 2014

5-87

CHAPTER 5 - SUBSTRUCTURES

2. Design of Mechanically Stabilized Earth Walls and Reinforced


Soil Slopes, Volumes I and II, November 2009, FHWA-NHI-10024 and FHWA-NHI-10-025
3. Corrosion/Degradation of Soil Reinforcements for Mechanically
Stabilized Earth Walls and Reinforced Soil Slopes, November
2009, FHWA-NHI-09-087
4. Standard Specification Section 636 Mechanically Stabilized
Earth Retaining Wall
It is the responsibility of the Geotechnical Designer to assess the wall for
bearing resistance, settlement, and global slope stability.
The calculation of lateral earth pressure on MSE walls should be as
specified in AASHTO LRFD 3.11.5.8.
MSE walls with steel reinforcement and precast panels are relatively low
in cost. These walls do require a high quality backfill with strict
electrochemical requirements, as defined in the Standard Specifications
Section 636 - Mechanically Stabilized Earth Retaining Wall. The base
width of MSE walls is typically 70% of the wall height, which requires
considerable excavation in a cut situation. Therefore, in a cut situation,
base width requirements usually make MSE structures uneconomical
and difficult to construct. It is best to limit the height to approximately 35
feet for routine projects.
Facing options depend on the aesthetic and structural needs of the wall
system. Facing options typically include precast modular panels with
various shapes and texturing options. The facing type used can affect
the ability of the wall to tolerate settlement, depending on whether
continuous vertical joints between adjacent panels are specified. Refer
to Section 5.6.1.8 Aesthetics for further guidance.
MSE walls are inherently flexible and can tolerate moderate settlements
without suffering structural damage, depending upon the MSE wall panel
shape and alignment.
MSE walls are not appropriate if very weak soils are present that will not
support the wall and that are too deep to be over excavated, or if a deep
failure surface is present that result in slope instability. In these cases, a
deep foundation or soil modification may be considered.
MSE walls may be used to retain soil supporting bridge substructures.
The substructure units may be either spread footings or pile supported,
with the following additional design criteria:

March 2014

5-88

CHAPTER 5 - SUBSTRUCTURES

o The MSE wall shall be designed to provide a service life of not


less than 100 years.
o For the analysis of spread footings on top of the reinforced soil
zone, a factored bearing resistance of 7 ksf should be used for
the strength limit state, and a factored bearing resistance of 4 ksf
should be used to limit settlements to less than approximately 0.5
inch.
o A minimum distance of 4 feet should be provided between the
bottom of the superstructure and the berm in front of the abutment
breastwall or pile cap and behind the MSE top panel, for future
bridge inspection and maintenance purposes.
o The minimum distance from the centerline of the bearing on the
bearing on the abutment to the outer edge of the MSE wall facing
should be 3.5 feet.
o A minimum distance of 2 feet should be provided between the
back of MSE wall panels and the front face of abutment or pile
cap.
o If the abutment is supported on piles or piles installed in sleeves,
a minimum distance of 2 feet should be provided to allow
compaction equipment to be used between piles or sleeves and
the back face of panels, and to allow a 15 reinforcing strap skew
to clear a typical 2-ft diameter pile sleeve.
o The top of the MSE panel in front of footings or pile caps should
be set 1 foot above the berm elevation.
o If embedding spread footings for frost protection within the
reinforced mass is impractical, provide at least 2 feet of soil cover
and place the footing on a minimum 3-foot thick bed of compacted
coarse aggregate.
o An impervious geomembrane consisting of low-permeability, 2sided textured HDPE a minimum of 60 mils thick should be
installed near the top of the reinforced soil zone to reduce the
chance of water and salt-laden water infiltration into the reinforced
backfill. The membrane should be bonded to the back face of the
abutment, and sloped to shed water that infiltrates from the road
surface.
o The need for fencing along the top of the wall should be
investigated on a project by project basis.

March 2014

5-89

CHAPTER 5 - SUBSTRUCTURES

Prior to selection of MSE walls for a project, consideration should be


given to the location of any utility behind or within the reinforced soil
backfill zone. It is best not to place utilities within the reinforced backfill
zone because it would be impossible to access the utility from the
ground surface without cutting through the soil reinforcement layers,
thereby compromising the integrity of the wall. Coordination of the wall
with project elements (such as drainage, utilities, luminaries, guardrail, or
bridge elements) is critical to avoid costly change orders during
construction. Moreover, failure of a sewer or water main located within
an MSE wall mass could result in failure of the wall. As a result, MSE
walls must not be used in areas where water and/or sewer utilities are
present. It is also best to locate drainage features and signal or sign
foundations outside of the MSE reinforced backfill zone.
Since MSE walls are proprietary and the wall vendor performs the
design, it is imperative that the design requirements be clearly stated on
the plans. If there are any unusual aesthetic requirements, design
acceptance requirements, or loading conditions for which the wall needs
to be designed, they should be clearly shown on the plans. Refer to
Section 5.6.5.5 PS&E for Project with Proprietary Walls for plan
development requirements.
MSE walls are measured and paid for by the area of wall face, as
determined from the approved shop drawings. The high quality backfill
and wall design are included in the MSE wall pay item. The Designer
should consider this when comparing the cost of MSE walls with other
wall systems, which typically pay for backfill as a separate pay item.
Excavation is also paid for separately as common excavation. The
Designer should consult the current Special Provision for measurement
and payment information.
B. Geosynthetic-Reinforced Soil Walls
Geosynthetic-reinforced soil (GRS) walls are MSE or Precast Gravity
Block walls with geosynthetic (polymeric) soil reinforcement. GRS walls
are designed to create a reinforced soil volume behind a wall facing.
Facing options include precast concrete modular panels or modular
concrete blocks. Geosynthetic facings, although available, are not
acceptable for permanent facing due to potential facing degradation
when exposed to sunlight. Facings consisting of dry-cast concrete are
susceptible to degradation caused by freeze-thaw and are not allowed.
GRS walls are not permitted in waterways.
GRS walls are constructed from the bottom up and are therefore best
suited for fill situations. The base width of GRS walls is typically 70% of
the wall height, which requires considerable excavation in a cut situation.

March 2014

5-90

CHAPTER 5 - SUBSTRUCTURES

It is best to limit the height of GRS walls to 20 feet or less for routine
projects.
GRS walls have a low cost and can handle significant settlement.
Compared to steel-reinforced systems, internal wall deformations may
be greater and electrochemical backfill requirements less strict, but a
high quality backfill is still required. Only geosynthetic products for which
long-term product durability is well defined per LRFD 11.10.6.4 will be
allowed.
GRS walls are proprietary and are designed by a wall manufacturer for
internal and external stability. GRS walls shall be designed with a
service life of not less than 75 years. The walls shall be designed in
accordance with the following:
1. LRFD Article 11.10
2. Design of Mechanically Stabilized Earth Walls and Reinforced
Soil Slopes, Volumes I and II, November 2009, FHWA-NHI-10024 and FHWA-NHI-10-025
3. Corrosion/Degradation of Soil Reinforcements for Mechanically
Stabilized Earth Walls and Reinforced Soil Slopes, November
2009, FHWA-NHI-00-087
It is the responsibility of the Geotechnical Designer to assess the wall for
bearing capacity, settlement, and global slope stability.
Since these preapproved walls are proprietary and the wall vendor
performs the design, it is imperative that the design requirements for
GRS wall be clearly stated on the plans. If there are any unusual
aesthetic requirements, design acceptance requirements, or loading
conditions or pressures for which the wall needs to be designed, they
should be clearly shown on the plans. Refer to Section 5.6.5.5 PS&E for
Project with Proprietary Walls for plan development requirements.
Coordination of the wall with project elements (such as drainage, utilities,
luminaries, guardrail, or bridge elements) is critical to avoid costly
change orders during construction. It is best to locate drainage
structures and signal or sign foundations outside of the reinforced
backfill zone.
5.6.5.5

PS&E for Project with Proprietary Walls

The PS&E package for a bridge project including proprietary wall item will
include the following:

March 2014

5-91

CHAPTER 5 - SUBSTRUCTURES

General wall plan

Wall profile, showing neat line top and bottom of the wall and
final ground line in front of and in back of the wall

Profiles showing the existing and final grades

Typical wall cross section with generic details including batter

Factored bearing resistance

Foundation embedment criteria

Leveling pad details

General details for any desired appurtenances, such as coping


or drainage requirements

Project specific loads for other design acceptance requirements


(examples: seismic loads, earth loads due to thermal movement
of abutments)

Special facing treatment (shape, texturing, color)

Project-specific construction requirements (example: crushed


stone)

Highway approach cross sections showing only the face of the


wall and footing

5.6.6 Geosynthetic Reinforced Soil Integrated Bridge Systems


GRS walls associated with Geosynthetic Reinforced Soil Integrated Bridge
Systems (GRS-IBS) designed in accordance GRS-IBS Interim Implementation
Guide, FHWA-HR-11-026, January 2011, may be considered for some
bridges over waterways, with the approval of the Assistant Bridge Program
Manager.
5.6.7 Anchored Wall Systems
5.6.7.1

CON/SPAN Wingwall

CON/SPAN wingwall systems may only be used in conjunction with


CON/SPAN precast drainage structures. The system consists of a
precast face panel with a precast concrete soil anchor located near the
base of the face panel. The wingwall system is connected to the
CON/SPAN drainage structure. The wall should be backfilled with
March 2014

5-92

CHAPTER 5 - SUBSTRUCTURES

granular borrow material suitable for underwater backfill and compacted


per the Standard Specifications. The maximum wall height available is
16.5 feet, and should only be used with a level backfill surface and seismic
loads less than a = 0.1g when a seismic analysis is required for design
(ASCE, 2001). Refer to Section 3.7.2 Seismic Analysis for guidance.
The CON/SPAN wingwall system should be designed in accordance with
LRFD 11.9 Anchored Walls. The design requirements for the
CON/SPAN wingwall system should be included with the contract
documents in Special Provision 534.
CON/SPAN wingwall system should be placed on a footing, which serves
both as a leveling slab and a structural foundation. This may include, but
is not limited to a cast-in-place concrete footing, cast-in-place stub wall
with footing, or a precast concrete footing meeting the requirements of
Section 5.2.1 Frost, Section 5.3 Spread Footings, and Section 2.3.11
Scour. The footing should be sized to support the weight of the wall
panels and weight of soil in and above the anchor system (ASCE, 2001).
The CON/SPAN wingwall system should be equipped with a drainage
system, consisting of a perforated drainage pipe installed in the backfill
behind the wall, which outlets through a 4 inch diameter weep hole cast in
the facing panel, per the manufacturers requirements (ASCE, 2001).
5.6.7.2

Metal Structural Plate Headwall/Wingwall

Metal structural plate headwall/wingwall may only be used in conjunction


with metal structural plate box culverts. However, preference should be
given to the use of a PCMG wall system for increased durability. The
headwall system consists of a metal structural plate face, which is
connected to the top of the metal structural plate box with an anchor rod.
The wingwall system consists of a metal structural plate face with a
deadman connected to the face with a tie rod and whale system. The
maximum wall height available is 14.25 feet.
The metal structural plate headwall/wingwall system should be designed
in accordance with the most recent version of AASHTO LRFD. The
design requirements for the metal structural plate headwall/wingwall
system should be included with the contract documents.
5.6.8 Gabions
Gabion walls consist of stacked 3 feet cubed wire baskets, which are filled with
stone. Groups of filled gabion baskets are stacked to construct a gravity wall.
Gabion walls should be designed as specified in Section 3.6.7.2 Prefabricated
Modular Walls. In designing gabion walls, a unit weight, , of 100 lb/ft3 should

March 2014

5-93

CHAPTER 5 - SUBSTRUCTURES

be used for the weight of stone inside the baskets. Gabion walls should be
backfilled with granular or gravel borrow. An angle of wall friction, , of 24
should be used for design. Wire for gabion baskets should be either PVCcoated or galvanized. A PVC coating is preferred as it does not flake off.
MaineDOT experience has shown that constructing gabion walls correctly can
be costly and time-consuming. Disadvantages in the use of gabions include
subjection to corrosion when placed in water and occurrence of vandalism by
the cutting of the basket wires. Gabion walls should be used only in noncritical situations, in dry environments, and in rural areas, where the probability
of corrosion and vandalism are less (MaineDOT, 2002). Gabion wall heights
in excess of 6 feet are not recommended.
5.7

Piles

5.7.1 General
Piles should be considered when spread footings cannot be founded on
bedrock or on competent soils at a reasonable cost. Piles should also be
considered where soil conditions permit use of spread footings, but where the
soils are susceptible to scour, liquefaction or lateral spreading.
Pile foundations should be designed so that the available factored
geotechnical and drivability resistance is greater than the factored loads
applied to the pile at the strength limit state. Service limit state design of
driven pile foundations includes an evaluation of settlement, overall stability,
lateral squeeze and lateral movement.
5.7.2 H-Piles
H-Piles used for bridge foundations should be comprised of rolled-steel
sections of ASTM A572, Grade 50 steel, with a minimum yield stress of 50 ksi.
Refer to Section 7.2.1 Structural Steel for H-pile material requirements.
5.7.2.1

Axial Resistance

The maximum factored axial design load applied to H-pile sections should
not exceed the lesser of the factored structural pile resistance, the
factored geotechnical pile resistance and the factored drivability
resistance. The factored structural resistance of H-pile sections should be
determined using a resistance factor, , listed below:

March 2014

c = 0.50 for axial resistance of piles in compression and


subject to damage due to severe driving where use of a pile tip
is necessary.

5-94

CHAPTER 5 - SUBSTRUCTURES

c = 0.60 for axial resistance of piles in compression under


good driving conditions, where use of a driving tip is not
necessary.

For combined axial and flexural resistance of undamaged pile, the


resistance factors are listed below:

c = 0.70 for axial resistance of H-piles in compression.

f = 1.00 for flexural resistance of H-piles.

The resistance factors, c and f, are to be used in interaction equations


in LRFD 6.9.2.2.
The factored axial structural axial resistances of selected H-Pile sections
are presented in Table 5-7. For the purposes of Table 5-7, the H-piles
were assumed fully braced, and an effective length factor (K) of 1.0was
used. The Structural Designer should recalculate structural resistances
for the upper and lower portions of the H-pile based on unbraced lengths
and K-values from project specific LPILE analyses and recalculate
structural resistances. For preliminary design purposes, however, the
resistances provided in Table 5-7 may be used to estimate the factored
structural axial resistance of that portion of the pile which is theoretically in
pure compression, i.e., that portion below the point of fixity.
Commentary: Experience in using 50 ksi steel for H-Pile foundations has
shown that the factored axial geotechnical resistance frequently governs
design. This is particularly apparent for end-bearing piles on poor-quality
and/or soft bedrock and for friction piles.

March 2014

5-95

CHAPTER 5 - SUBSTRUCTURES

Table 5-7 Factored Axial Structural Resistance of Selected H-Pile


Sections
Fy = 50 ksi and fully braced
Pile Section

HP 10x42+
HP 10x57
HP 12x53+
HP 12x63
HP 12x74
HP 12x84
HP 14x73+
HP 14x89+
HP 14x102
HP 14x117

Factored Axial Structural Resistance


Good driving
Severe driving
conditions
conditions
= 0.60
= 0.50
(kips)
(kips)
372
310
504
420
465
388
552
460
654
545
738
615
642
535
783
653
900
750
1032
860

Note: Those marked + are preferred sections


The factored geotechnical and drivability resistances should be
determined for site-specific conditions by the Geotechnical Designer.
Consideration should be given to downdrag, soil relaxation, soil setup,
lateral spreading and any other site-specific factors, which may affect the
pile capacity during and after construction. The factored geotechnical
resistance should be determined by applying a resistance, factor which is
dependent on the design method.
5.7.2.2

Lateral Pile Resistance for the Service Limit State

Horizontal movement of pile groups induced by lateral loads shall be


evaluated for Service Limit State Design. The lateral resistance of a pile is
governed by the loading condition, pile stiffness, stiffness of the soil, and
the degree of fixity. The lateral resistance (PL) and depth to fixity (Df), for
service limit state design for selected H-Pile sections in sand and clay are
presented in Table 5-8 and Table 5-9, respectively. The factored lateral
resistances presented in Tables 5-8 and 5-9 assume a resistance factor of
1.0 and a maximum lateral deflection of 1/8 inch.

March 2014

5-96

CHAPTER 5 - SUBSTRUCTURES
Commentary: The lateral resistance and depth to fixity presented in
Tables 5-8 and Table 5-9 were determined using the computer program
LPILE Plus Version 4, the soil properties stated, a fixed condition at
the pile head, an infinitely long pile, an applied axial load equal to As x
0.25 x Fy and a deflection of 1/8.

Table 5-8 Factored Lateral Resistance and Depth to Fixity for Strength
Limit State Design for H-Pile Sections in Sand, =1.0
Loose
Pile
Section
HP 10x42+
HP 10x57
HP 12x53+
HP 12x63
HP 12x74
HP 13x60
HP 13x73
HP 13x87
HP 14x73+
HP 14x89+
HP 14x102
HP 14x117

PL
(kips)
6.2
7.1
8.1
8.9
9.4
9.0
9.8
10.6
10.5
11.4
12.3
13.1

Df
(ft)
24
26
28
30
31
31
32
32
32
33
35
36

Medium Dense
PL
Df
(kips)
(ft)
9.9
20
11.4
22
13.3
24
14.4
25
15.6
25
15.0
25
16.4
26
17.7
26
17.8
26
19.5
27
20.9
28
22.3
29

Dense
PL
(kips)
11.7
13.6
16.1
17.4
18.9
18.2
20.0
21.7
21.9
24.1
25.9
27.0

Df
(ft)
18
19
20
21
22
21
22
23
23
24
25
25

Note: Those marked + are preferred sections. PL and Df are


determined assuming a friction angle, , of 32.
Where the applied lateral load from the Service Limit State Load Combination
exceeds that presented in Tables 5-8 and 5-9, or the pile length is less than the
depth to fixity shown in the table, a more thorough analysis is recommended,
using actual loading and soil conditions. Where soils differ from the conditions
assumed in the tables, the Designer should complete a more thorough analysis.
Tables 5-8 and 5-9 present the lateral resistance and depth to fixity for a lateral
load applied perpendicular to the pile flange. For conventional abutments and
mass piers, H-piles should be oriented with the flange perpendicular to the
substructure axis in the direction of the maximum applied lateral load. For
conventional abutments and mass piers, where H-piles are oriented with the web
perpendicular to the maximum applied lateral load, a thorough analysis of the
foundation is recommended, using actual loading and soil conditions (Tables 5-8
and 5-9 do not apply). For integral abutments where the web is oriented
perpendicular to the principal axis, the design should be in accordance with
Section 5.4.2 Integral Abutments.

March 2014

5-97

CHAPTER 5 - SUBSTRUCTURES

Table 5-9 Factored Lateral Resistance and Depth to Fixity for


Service Limit State Design for H-Pile Sections in Clay, =1.0, Load
Perpendicular to Flange
Soft1
Pile
Section
HP 10x42+
HP 10x57
HP 12x53+
HP 12x63
HP 12x74
HP 13x60
HP 13x73
HP 13x87
HP 14x73+
HP 14x89+
HP 14x102
HP 14x117

PL
(kips)
5.1
5.5
6.3
6.7
7.1
7.0
7.5
7.9
8.1
8.7
9.1
9.5

Df
(ft)
22
24
26
27
27
27
28
29
29
31
31
32

Medium Stiff2
PL
Df
(kips)
(ft)
9.2
18
10.2
20
11.7
21
12.4
22
13.1
22
12.8
22
13.8
23
15.6
25
14.8
24
15.9
25
16.7
26
17.5
26

Stiff3
PL
(kips)
13.1
14.5
16.6
17.6
18.7
18.2
19.5
20.7
21.0
22.5
23.6
24.8

Df
(ft)
16
18
19
19
20
19
21
21
21
22
22
24

Note: Those marked + are preferred sections.


1
Su = 375 psf, 2Su = 750 psf, 3Su = 1125 psf

March 2014

5-98

CHAPTER 5 - SUBSTRUCTURES

5.7.3 Layout and Construction


The pile spacing should not be larger than is reasonable or practical. The
center-to-center pile spacing should not be less than 30 inches or 2.5 to 3
times the pile diameter. A reasonable maximum spacing for piles in the back
row of abutments is 12 feet.
Care should be exercised in locating piles to avoid interference with other
piles, both in the final position and during the driving process. If a plumb pile
in the back row is located directly behind a battered pile in the front row, the
Contractor may be forced to plan his sequence of pile driving and cut-offs in a
less efficient manner than if the back row of piles were staggered with the front
row.
The distance from the side of any pile to the nearest edge of the pile cap shall
not be less than 9.0 inches. The tops of piles should project at least 18 inches
into the pile cap after all damaged pile material has been removed.
All piles should be equipped with a driving shoe. Refer to Standard
Specification Section 501 Foundation Piles for further guidance.
5.7.4 Concrete Piles
Concrete piles are used as displacement piles provided they can be driven
without damage, that is, there are no boulders or hard driving dense soils.
Two types of concrete piles are precast conventionally reinforced and precast
prestressed piles. Both types are of constant cross section, though they may
have tapered tips. Pile shapes include square, octagonal, and round sections
and may be either solid or hollow. Typical pile cross sections used range from
10 inches to 16 inches, but sizes above and below this range are also
produced. Refer to LRFD Article 5.13.4, Concrete Piles, and FHWA, 1998 for
detailed information regarding concrete piles.
Precast concrete piles are suitable for use as friction piles when driven in
sand, gravel, or clays. Precast concrete piles are capable of high capacities
when used as end bearing piles. In boulder conditions, a short piece of
structural H-pile section or stinger may be cast into or attached to the pile for
penetration through the zone of cobbles and boulders.
Conventionally reinforced concrete piles (concrete with reinforcing steel bars
and spiral reinforcing steel cages) are susceptible to damage by mishandling
or driving. Prestressed concrete piles are more vulnerable to damage from
striking hard layers of soil or obstructions during driving than conventionally
reinforced concrete piles. Piles should be equipped with a metal driving shoe
for hard driving conditions. High stresses during driving can cause cracking in
all concrete piles.
March 2014

5-99

CHAPTER 5 - SUBSTRUCTURES

Precast piles are difficult to splice, particularly prestressed piles. Accurate


knowledge of pile lengths is required when using concrete piles, as they are
also difficult to shorten. Special precautions should be taken when placing
concrete piles during cold weather. Temperature gradients can cause
concrete to crack due to non-uniform shrinkage and expansion.
A concrete pile foundation design should consider that deterioration of
concrete piles can occur due to sulfates in soil, ground water, or sea water;
chlorides in soils and chemical wastes; or acidic ground water and organic
acids. Laboratory testing of soil and ground water samples for sulfates and pH
is usually sufficient to assess pile deterioration potential. A full chemical
analysis of soil and ground water samples is recommended when chemical
wastes are suspected.
5.7.5 Steel Pipe Piles
5.7.5.1

Design - General

The maximum factored applied axial load on any pipe pile shall not
exceed the lesser of the factored structural compressive resistance, the
factored axial geotechnical resistance and the factored drivability pile
resistance. For the strength limit state, the factored axial compressive
structural resistance of pipe piles (Pr) shall be estimated using the
following resistance factors (c):

c = 0.60 for piles subject to damage in severe driving


conditions where use of a pile tip is necessary

c = 0.70 for piles under good driving conditions where use


of a pile tip is not necessary

The nominal compressive structural resistance (Pn) for pipe piles loaded in
compression should be estimated as specified in LRFD 6.9.5.1 using the
column slenderness factor, .
At the strength limit state an axial resistance factor, c, of 0.80, and a
flexural resistance factor, f, of 1.0 should be applied to combined
nominal axial and flexural resistance in the interaction equation in LRFD
6.9.2.2.
5.7.5.2

Material and Design Section

Pipe piles consist of seamless, straight butt-welded or spiral butt-welded


metal shells. Steel pipe piles may be driven in groups, to support groundlevel pile caps, or in-line to form pile bents. They are available in a wide
range of diameters. Typical wall thicknesses are limited to the range of

March 2014

5-100

CHAPTER 5 - SUBSTRUCTURES

1/2 to 1 inch. MaineDOT practice has commonly limited their use to 24 to


32 inch diameters when used in pier bents. All pipe piles are filled with
Class A concrete after driving. Additionally, pipe piles employed as pier
bents are internally reinforced with a reinforcing cage.
Concrete filled pipe piles have a high load-carrying capacity and provide
high bending resistance where an unsupported length is subject to lateral
loads. For design criteria and corrosion protection of pipe piles in pier
bents, refer to Section 5.5.2.5 Pile Protection and 5.5.2.6 Pipe Pile
Coatings and Cathodic Protection.
Pipe piles may be driven open or closed ended. If the capacity from the
full pile toe is required, the pile should be driven closed ended, with a flat
plate or conical tip. Closed ended types are preferred, except if the pile is
designed as a friction displacement pile.
If obstructions are expected, the pile should be open-ended, so that it can
be cleaned out and driven further. Open-ended piles driven in sands or
clays will form a soil plug at some stage during driving. At this stage, the
pile acts like a closed ended pile and can significantly increase the pile toe
resistance. Piles driven open-ended should be cleaned, leaving a length
of soil plug ranging from two to three pile diameters, and filled with
concrete after driving.
Steel pile material should conform to ASTM A252 Grade 3. Open-ended
piles should be reinforced with steel cutting shoes to provide protection
against damage. When pipe piles are driven to weathered bedrock or
though boulders, an end plate or conical point with a rounded nose is
often used to prevent distortion of the pile nose. End closures should be
cast steel, conforming to the requirements of ASTM A27 (grade 65-35) or
ASTM A148 (grade 90-60).
For high vertical or lateral loads, open-ended pipe piles may be socketed
in bedrock. They can also have a structural shape such as an H-section
inserted into the concrete and socked into bedrock. Anchoring pipe piles
with rock dowels or anchors is not recommended and should only be
considered when the preceding alternatives are found to be not feasible.
Pipe piles can be spliced using full penetration groove welds or proprietary
splicing sleeves that provide full strength in bending.
5.7.6 Downdrag
Where the soil deposit in which piles are installed is subject to settlement,
downdrag forces may be induced on piles. As little as 1/2 of differential
settlement may induce downdrag forces. Downdrag loads reduce the usable

March 2014

5-101

CHAPTER 5 - SUBSTRUCTURES

pile capacity. Possible development of downdrag loads on piles should be


considered when:

Sites are underlain by compressible clays, silts, or peats

Fill has been recently placed on the surface

The groundwater has been substantially lowered

Downdrag loads should be considered as permanent additional axial loads


when the nominal bearing capacity of the pile foundation is evaluated, and
when settlement of the pile foundation is evaluated.
To calculate downdrag loads on piles, the traditional approach is the total
stress -method, which is used for computing downdrag in cohesive soils.
Newer methods are based on the relationship between pile movement and
negative shaft resistance, and described in Briaud and Tucker (1993). The
downdrag loads should be factored by the appropriate load factor for
downdrag, p, and added to the factored vertical dead load applied to the pile.
If downdrag forces are significant, they can be reduced by applying a thin coat
of bitumen of the pile surface (Dixon, et. al., May 1998). Battered piles should
be avoided where downdrag loads are expected due to induced bending
moments in response to settlement. These bending moments can result in
pile deformation. In situations where downdrag forces cannot be reduced by
applying bitumen coating, the Designer should consider:

Forcing soil settlement prior to driving piles by preloading and


consolidation the soils

Using lightweight fills

Increasing the pile size

Sleeve piles

5.7.7 Pile Installation Quality Control and Nominal Pile Resistance


The nominal resistance a pile is driven to in the field is a function of the level
of quality assurance/control provided during construction operations. The
resistance factors for nominal pile resistances are presented in Table 5-10.
These resistance factors are based upon construction quality control beyond
the standard subsurface exploration and static pile capacity analysis.

March 2014

5-102

CHAPTER 5 - SUBSTRUCTURES

Table 5-10 Resistance Factors for Driven Piles


Construction Control Method
Static load test of at least one pile, with dynamic testing of
at least 2% to 5% of the production piles.
Dynamic testing with signal matching of at least 1 pile per
substructure, but no less than 2 dynamic pile tests from
opposite corners for substructures longer than 40 feet or
with more than 15 piles, but no less than 2% of the
production piles at any one site, and up to 5% of the
production piles for sites with moderate to highly variable
subsurface conditions.
Wave equation analysis without dynamic measurements or
load test

Resistance
Factor,
dyn
0.80
0.65

0.40

A pile group is classified as nonredundant if there are less than five (5) piles in
the group. If a pile group is nonredundant, past LRFD practice dictated a 20
percent reduction of the pile resistance factors, dyn, provided in Table 5-10,
and should be considered to provide a uniform level of safety.
Pile testing programs should include, at a minimum, wave equation analyses.
Wave equation analyses confirm that the design pile section can be installed
to the desired depth and ultimate capacity, without exceeding allowable pile
driving stresses, with an appropriate driving system and criteria.
In addition to wave equation analyses, pile testing programs should also
include dynamic load tests or, rarely, static load tests. Dynamic testing with
signal matching should be considered in order to:

Field-verify the nominal pile axial resistance

Establish driving criteria

Monitor piles installed in difficult subsurface conditions, such as soils


with obstructions and boulders, or a steeply sloping bedrock surface

Verify consistent hammer operation during extended pile installation


operations

Justify higher resistance factors

In general, the pile testing program should be commensurate with the design
assumptions; for example, at least 1 pile per bearing stratum will be tested.

March 2014

5-103

CHAPTER 5 - SUBSTRUCTURES

Pile testing programs should specify the number, location, and time of all
dynamic tests and/or static pile tests. When a dynamic load test program is
specified, the following requirements shall apply:

For large pile groups with more than 20 piles, the first and second
pile tests shall be conducted at opposite corners of the substructure,
and at least one additional dynamic test shall be conducted midproduction, after approximately one half of the production piles have
been installed.

Post-driving analyses (CAPWAP) are required.

Provisions for 24 to 72 hour pile restrikes shall be included, for


substructures where setup or relaxation effects are expected.

Provisions for 24 to 72 hour dynamic restrike tests are mandatory for


friction piles or piles designed to end bear in any strata other than
bedrock.

Provisions should be provided for the conduct of additional dynamic


load tests during production, for field verification that the driving
criteria are consistently achieving the required nominal pile
resistances.

A minimum of 2% of the piles shall be tested when dynamic (or static) testing
is specified. It may be necessary to test 5% or more piles, when there are
more than 20 piles in a substructure, when difficult driving is expected, when
variable or inconsistent soil conditions are expected, or when additional tests
during production are necessary to verify hammer performance and
geotechnical resistances.
The establishment of the driving criteria should include limiting driving stresses
to the following thresholds:

For steel piles in compression and tension, driving stresses should


not exceed 90% of the yield strength of the pile material. For 50 ksi
steel, this results in a maximum driving stress of 45 ksi.

For concrete filled pipe piles, if unfilled when driven, driving stresses
should not exceed 90% of the yield strength of the steel shell
material.

For concrete piles, driving compressive stresses should not exceed


0.85 times the concrete compressive strength. Tensile stresses
during driving should not exceed 0.70 times the yield strength of the
steel reinforcement.

March 2014

5-104

CHAPTER 5 - SUBSTRUCTURES

5.8

For prestressed concrete piles, driving compressive stresses should


not exceed 0.85 times the concrete compressive strength minus the
effective prestress. Tensile stresses during driving should be limited
to 0.095 times the square root of the compressive strength (ksi) plus
the effective prestress.

Drilled Shafts

Drilled shafts may be an economical alternative to spread footings or pile


foundations. Drilled shafts can be an advantageous foundation alternative when:

Spread footings cannot be founded on suitable soil, or bedrock, within a


reasonable depth or when driven piles are not viable.

Traditional piles would result in insufficient embedment depth and rocksocketed deep foundations are needed.

Scour depth is large.

Foundations are required in stream channels. Drilled shafts will avoid


expensive construction of cofferdams. Advantages are the reduction of
the quantities and cost of excavating, dewatering, and sheeting, and in
limiting environmental impact.

The elimination of waterline footings is advantageous and possible by


extending drilled shafts as a column up to the pier cap.

The foundation is required to resist high lateral loads or uplift loads.

There is little tolerance for deformation.

The cost and constructability of seals and caps for pile supported
structures is high.

Although there are many references for the design and analysis of drilled shafts,
MaineDOT follows the procedures found in FHWA, 2010 and LRFD Article 10.8.
The structural design of drilled shafts is similar to the LRFD method for a column
with axial load and bending, and shear. Interaction diagrams should be
developed to assess resistance to combined axial and bending.
The Bridge Program has developed a Special Provision to govern the
construction of drilled shafts. Consult the Geotechnical Designer for the current
version.

March 2014

5-105

CHAPTER 5 - SUBSTRUCTURES

5.9

Embankment Issues

Embankment design considerations include settlement, slope stability, and


bearing capacity at the base. Special design requirements for embankments will
be presented in the Geotechnical Report. The Geotechnical Designer should
review plans to determine any special design requirements with regard to an
embankment.
5.9.1 Embankment Settlement
The embankment settlement should be evaluated using the methods
discussed in Section 5.3.6 Settlement and must be within tolerable limits.
Differential settlement is more of a concern than total settlement and should
be evaluated by the Geotechnical Designer. Tolerable settlement also
depends upon the structural integrity of the bridge or culvert and should be
coordinated with the Structural Designer.
If settlement exceeds the tolerable limits, or the time needed to allow for
settlement is excessive, several methods to address this are available to the
Designer:

Compressible materials can be removed and replaced to limit


settlements.

Preloads alone or in combination with surcharge can be used to


complete settlements prior to construction.

Prefabricated vertical drains can be used in conjunction with preloads


to accelerate settlements.

Lightweight fill materials such as tire shreds, geofoam or light weight


concrete fill can be used.

The use of a preload, surcharge, or prefabricated vertical drains should be


accompanied by the use of instrumentation (settlement platforms,
piezometers, inclinometers) to assist in determining that an acceptable level of
consolidation has taken place.
5.9.2 Embankment Stability
Embankment stability problems most often occur where embankments are to
be built over soft weak soils such as low strength clays, silt, or peats. There
are three major types of instability that should be considered in the design of
embankments over weak foundation soils: circular arc failure, sliding block
failure, and lateral squeeze. These stability problems are defined as external
stability problems. Internal stability problems generally result from the

March 2014

5-106

CHAPTER 5 - SUBSTRUCTURES

selection of poor quality materials and/or improper placement requirements.


Refer to Section 5.3.7 Overall Stability for methods of analysis.
Once the soil profile, soil strengths, and depth of water table have been
determined by both field explorations and field and laboratory testing, the
stability of the embankment can be analyzed. The evaluation of slope stability
of earth slopes with or without a foundation unit should be investigated at the
Service I Load Combination and an appropriate resistance factor. The
resistance factor, , may be taken as:

0.75 - where the geotechnical parameters are well defined, and the
slope does not support or contain a structural element

0.65 where the geotechnical parameters are based on limited


information or the slope contains or supports a structural element.

Available slope stability programs produce a single factor of safety. In light of


this, the past practice of checking overall slope stability using ASD methods
may be continued to insure that slopes and slopes with footings have a factor
of safety equivalent to 1.3 and 1.5, respectively.
If the load and resistance balance cannot be met, several methods to improve
stability can be undertaken:

Removal and replacement of the weak material

Use of a mid-slope berm or other variations of berms

Soil reinforcement with steel, geogrid, or geotextile

Installation of prefabricated vertical (wick) drains, sand drains, or


stone columns

Instrumentation and control of embankment construction

Installation of a structural support such as a retaining wall

Lateral squeeze can occur when the lateral movement (consolidation) of soft
soils transmits an excessive lateral thrust, which may bend or push an
adjacent substructure. The best way to minimize lateral squeeze is to
complete embankment settlements prior to construction of adjacent
substructures.

March 2014

5-107

CHAPTER 5 - SUBSTRUCTURES

5.9.3 Embankment Bearing Capacity


The embankment bearing resistance should be evaluated using the methods
discussed in Section 5.3.5 Bearing Resistance. The factored bearing
resistance should equal or exceed the factor applied loads.
5.9.4 Embankment Seismic Considerations
Currently, there are no LRFD codes for embankment seismic design.
Therefore, using allowable stress design methods, a minimum seismic factor
of safety of 1.0 is acceptable for slope stability and liquefaction. Refer to
Section 3.7.4 Embankment & Embankments Supporting Substructure Units.
Should poor seismic performance of an embankment impact the overall
serviceability or performance of a critical structure the Department may specify
a higher level of seismic performance or specify appropriate seismic
provisions.
If the seismic slope stability factor of safety falls below 1.0 using the seismic
coefficient-factor of safety method, a permanent seismic deformation analysis
should be conducted using the Newmark Method (Newmark, 1965). This
method approximates the cumulative vertical deformation or settlement at the
back of the slope for a given earthquake ground motion. The failure mass is
modeled as a block on a plane. A maximum allowable seismic settlement of 6
inches at a bridge approach, resulting from the design earthquake event, is
considered acceptable. Refer to Section 3.7 Seismic for loading
considerations.

March 2014

5-108

CHAPTER 5 - SUBSTRUCTURES

References
Abendroth, Greimann, 1989, Rational Design Approach for Integral Abutment
Bridge Piles, TRR 1223
AASHTO, 2012, AASHTO LRFD Bridge Design Specifications, 6th Edition, with
2013 Interim Revisions, Washington, DC
AASHTO, 2011, AASHTO Guide Specifications for LRFD Seismic Bridge Design,
2nd Edition, with 2014 Interim Revisions, Washington, DC.
AASHTO, 2002, Standard Specifications for Highway Bridges, Seventeenth
Edition, Washington, DC
American Society of Civil Engineers, 2001, Evaluation of CON/SPAN Wingwall
System, Highway Innovative Technology Evaluation Center
Briaud and Tucker, 1993, Downdrag on Bitumen Coated Piles, NCHRP study
Burke, M., Gloyd, S., 1996, Semi-Integral Bridges: A Revelation, TRB Paper No.
960626, Transportation Research Board 1994 Annual Meeting, Washington, DC
COM624P, FHWA computer program
Dixon, L. A., and Sandford, T.C., 1998, Effects of Bitumen Coating on the Axial
and Lateral Loading of Abutment Piles Subject to Downdrag, MaineDOT Report
No. 95-04 by University of Maine, May
Delano, J.G.; Davids, W. G. ; Sandford, T. C.; Krusinski, L., 2005, Behavior of
Pile-Supported Abutments at Bridge Sites with Shallow Bedrock Phase I,
MaineDOT Technical Report No. ME 01-7, June
FB-Multipier, Bridge Software Institute computer program
Federal Highway Administration, 2006, Design and Construction of Driven Pile
Foundations, Volumes I and II, FHWA Publications NHI 05-042 and NHI-05-043,
PDCA
Federal Highway Administration, 2010, Drilled Shafts: Construction Procedures
and LRFD Design Methods, Publication No. FHWA-NHI-10-016
Federal Highway Administration, Evaluation of Soil and Rock Properties,
Geotechnical Engineering Circular (GEC) No. 5
Federal Highway Administration, 1980, Integral, No-Joint Structures and
Required Provisions for Movement, FHWA Technical Advisory, T140.13, January
28

March 2014

5-109

CHAPTER 5 - SUBSTRUCTURES

Federal Highway Administration, 2002, Shallow Foundations, Geotechnical


Engineering Circular (GEC) No. 6, FHWA-IF-02-054, September
Federal Highway Administration, 2011, Geosynthetic Reinforced Soil Integrated
Bridge System Interim Implementation Guide, FHWA-HRT-11-026, January.
Hartt, S.L, Sandford, T.C., Davids, W.G., 2006, Monitoring a Pile-Supported
Integral Abutment Bridge at a Site with Shallow Bedrock - Phase II, MaineDOT
Technical Report No. ME 01-7, August
LPILE, Ensoft Inc. computer program
MaineDOT, 2010, Maine Department of Transportation Pre-Approval Process for
Proprietary Wall Systems, January 2
MaineDOT, 2002, Value Engineering Report
National Corporate Highway Research Program, 1991, Manuals for the Design of
Bridge Foundations, Transportation Research Board, Report 343, Washington
DC
Newmark, N.M., 1965, Effects of Earthquakes on Dams and Embankments,
Geotechnique, Vol. 15
Massachusetts Highway Department, 1999, Mass Highway Bridge Manual
Sandford, T. C., 1994, Evaluation of Lateral Squeeze, MaineDOT Technical
Report 91-3 by University of Maine, October
Seed, H.B., Idriss, I. M., 1982, Ground Motions and Soil Liquefaction During
Earthquakes, Monograph No. 4, Earthquake Engineering Research Institute,
Berkeley, California, 134 p

March 2014

5-110

Potrebbero piacerti anche